c 2013 by nachiketa chakraborty. all rights reserved

169
c 2013 by Nachiketa Chakraborty. All rights reserved.

Upload: others

Post on 29-Dec-2021

2 views

Category:

Documents


0 download

TRANSCRIPT

c© 2013 by Nachiketa Chakraborty. All rights reserved.

STANDARD MODEL FOREGROUNDS IN NUCLEAR AND PARTICLEASTROPHYSICS

BY

NACHIKETA CHAKRABORTY

DISSERTATION

Submitted in partial fulfillment of the requirementsfor the degree of Doctor of Philosophy in Astronomy

in the Graduate College of theUniversity of Illinois at Urbana-Champaign, 2013

Urbana, Illinois

Doctoral Committee:

Professor B. D. Fields, ChairProfessor P. M. RickerAssociate Professor A. J. KemballProfessor J. C. Peng

Abstract

Nuclear and particle astrophysics comprises of applying standard principles of nuclear and

particle physics to astronomical systems and making predictions for observation. Also, ex-

treme environments in astronomical systems often serve as a probe for re-evaluating fun-

damental principles of physics and astronomy. The research described here comprises of

both these aspects of nuclear and particle astrophysics. There are two broad themes : pri-

mordial nucleosynthesis and cosmic gamma-ray sources. As a part of the first theme, the

“cosmological lithium problem”, and a possible nuclear physics solution to it is examined.

The theoretical prediction of the abundance of 7Li produced in the big bang exceeds the

observationally inferred abundance by a factor of 3-4, that is very difficult to explain within

the Standard Model of particle physics and cosmology. The possibility of missed nuclear

resonances in the network of nuclear reactions used to compute the theoretical primordial

abundance is discussed. Three such candidate resonances are found prompting experiments

to either find them or rule them out. In the light of recent experiments eliminating the most

promising of these candidates, requirement of new physics beyond the Standard Model looms

large as a a possible solution. In the second area of research, properties of cosmic gamma-ray

emitting sources such as star forming galaxies and blazars are studied. The gamma-ray sky

has contributions from these guaranteed sources and potentially others such as particle dark

matter. In the light of this, gamma-rays properties of the guaranteed astronomical sources

as spectral slope, photon count statistics and polarisation of X-rays and soft gamma-rays

is studied in the hope of achieving the broad goal of understanding the underlying physics

of radiative emission of these sources and hopefully disentangling these “foregrounds” from

ii

other exotic sources such as particle dark matter.

iii

Acknowledgments

My research and life in general throughout my Ph.D. would not have been successful without

the contribution of several people.

First and foremost, my advisor Brian Fields has and continues to be a tremendous in-

stitution for my training as a scientist. In a profession where the input of your mentor is

paramount, I have been fortunate to have a phenomenal scientist and a superlative person

to mould me into a scientist. I have learnt everything from scientific aspects of astrophysics

to the social aspects of building collaborations with other scientists from Brian. Alongside

doing good research, communicating it well is also very important. In this aspect, my com-

munication skills through talks and other means have improved significantly under Brian’s

tutelage. My entire research experience with Brian has been a thoroughly enjoyable expe-

rience. I am very, very lucky to have forged a strong, fruitful collaboration with a scientist

and person, I deeply admire. And I sure wish to continue working on interesting projects

with Brian throughout my career.

I have also been very fortunate to work with Prof. Athol Kemball. It was terrific fun to

do radio astronomy stuff with him, which brought back memories of my Master’s research.

My training with him in technical aspects of the field was also, in no small part responsible

for my successful application with my future employers. He has been extremely kind to me

and I hope to continue our collaboration in future.

Also, I have been privileged to have worked with two giants of particle physics and

astrophysics, Prof. Keith Olive and Prof. Scott Dodelson. My first paper with Prof. Olive

was a huge learning experience, and absolute delight and occasional fright to keep up with

iv

the lightning speed of thought and profound depth of understanding. Having read Prof.

Dodelson’s book on Modern Cosmology and papers, collaborating with him has been a

highlight of my Ph.D. work. Once, again it was a lot of fun in addition to being a huge

challenge to try to keep up with the high standards of research.

The support and encouragement of senior graduate students is very important to the

growth of a new graduate student. I have been very fortunate to have wonderful seniors who

have become great friends aside from being fine examples to look upto.

My friend and senior at UIUC, Rishi Khatri has been a huge influence and support even

before the time I reached UIUC. From helping me settle down in a new country to guiding

me about some research decisions with his characteristic candour, he has and continues to

be an inspiration and an excellent friend.

One of my most favourite persons who also belongs to the Fields research family and

has been a very good friend and a part of my research support system is Amy Lien. She

is without a doubt, a true protege of Brian, always being the nicest person and also an

excellent scientist. My collaboration and friendship with Amy has enriched my life as a

graduate student, and I cherish our friendship and collaboration.

Another member of the Fields research family, the prolific Vasiliki Pavlidou has been a

tremendous source of encouragement and inspiration. I have begun collaborating with Vaso

towards the end of my Ph.D. and I look forward to many papers together. I also, look

forward to her eminently enjoyable company as I head to Europe.

My life as a graduate student has been so very enjoyable due to the presence of good

friends in this very tightly knit department. Kuo-Chuan Pan, I-Jen, Hotaka, Yiran and

David have been wonderful friends and classmates. They have always been there for me

research and otherwise. I will always have fond memories of all the fun times we have had

together as a group and feel lucky to have made such good friends.

This department and its various members have been very kind and helpful to me at every

step of my six years of graduate life. Jeri, Mary-Margaret, Sandy, Bryan and Judy have been

v

extremely helpful and made it so much easier for me to do my research with everything else

taken care of. I certainly owe them a lot for comfortable life as a graduate student. And

Jeri, of course, is the kindest and sweetest person on this planet ! I will miss her a lot along

with everyone else in the department.

A huge part of my support system in Urbana-Champaign, are my friends from outside

the astronomy department. I cherish my friendship with Vineeth, Sarang, Pragnyadipta,

Sandeep, Vineet, Anika and the quintessential scientist Ritoban Basu Thakur. They have

all been of tremendous help in me settling down in the US and shared an unforgettable

common experience. My times spent taking classes and discussing science with Rito or

RBT, is something that has helped me strive towards doing exciting science. He has made

me better at being a scientist. Sandeep has been the most influential, my mentor at physical

training which has made me a lot sharper at my work. Indeed their friendship has been

invaluable and the gain of a lifetime that I take back with me.

Finally, my family including my parents, grandparents and my wife Soma, has been

my ultimate support system that made it possible for me to stay away from them and yet

function with some degree of success. My parents have been instrumental in me pursuing

science while being in India, still not a very popular choice. However, their atypical choices

have inspired me to do the same. Its their constant support and encouragement in chasing

my ambition of being a physicist that has given me freedom to do what I want, the way I

want. My father, Maloy who did his Ph.D. in Chemistry has always been a great source of

support, inspiration and encouragement taking a lot of detailed interest in my career. The

pride my parents take in my accomplishments is the biggest reward I gain.

My wife Soma has been the ideal partner in every possible way. It is testament to her

character, that she has never made us feel the geographical distance and has been central

to my growth as a scientist and person. She has helped me maintain the balance of the

personal and the professional in my life, impeccably. Being a researcher herself, Soma has

shown great patience and understanding throughout my Ph.D. and given a lot of importance

vi

to the accomplishment of my long and short term goals. At the same time, she has kept

me honest with my feet on the ground at all times. And therefore, when in doubt both

professionally and personally, she is the one I go to. I look forward to my future with her

in Europe. I hope, we can continue to support and improve each other as a scientist and a

person.

This work is a culmination of contributions from all these people, and particularly my

family. I dedicate my thesis to my family, particularly my beloved late grandfather who has

always had a lot of pride in my achievements.

vii

Table of Contents

Chapter 1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.1 Primordial nucleosynthesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.2 Cosmic gamma-rays . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71.3 High energy polarisation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111.4 Radio Supernovae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

Chapter 2 Resonant Destruction of Lithium During Big Bang Nucleosyn-thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162.2 Semi-analytical estimate of important reaction rates . . . . . . . . . . . . . 192.3 Systematic Search for Resonances . . . . . . . . . . . . . . . . . . . . . . . . 22

2.3.1 General Considerations . . . . . . . . . . . . . . . . . . . . . . . . . 222.3.2 List of Candidate Resonances . . . . . . . . . . . . . . . . . . . . . . 24

2.4 Narrow resonance solution space . . . . . . . . . . . . . . . . . . . . . . . . . 282.4.1 A = 8 Compound Nucleus . . . . . . . . . . . . . . . . . . . . . . . . 302.4.2 A = 9 Compound Nucleus . . . . . . . . . . . . . . . . . . . . . . . . 332.4.3 A = 10 Compound Nucleus . . . . . . . . . . . . . . . . . . . . . . . 372.4.4 A = 11 Compound Nucleus . . . . . . . . . . . . . . . . . . . . . . . 47

2.5 Reduced List of Candidate Resonances . . . . . . . . . . . . . . . . . . . . . 492.6 Discussion and Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

Chapter 3 Inverse-Compton Emission from Star-forming Galaxies . . . . . 553.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 563.2 Order-of-Magnitude Expectations . . . . . . . . . . . . . . . . . . . . . . . . 603.3 Targets: Interstellar Photon Fields . . . . . . . . . . . . . . . . . . . . . . . 633.4 Projectiles: Cosmic-Ray Electron Source and Propagation . . . . . . . . . . 673.5 Inverse Compton Emission from Individual Star-Forming Galaxies . . . . . 72

3.5.1 Inverse Compton Emissivity of a Star-Forming Galaxy . . . . . . . . 723.5.2 Total Inverse-Compton Luminosity from a Single Galaxy . . . . . . . 75

3.6 Results: Inverse Compton Contribution to the Extragalactic Background . . 803.7 Discussion and Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . 883.8 Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91

Chapter 4 Statistics of gamma-ray sources . . . . . . . . . . . . . . . . . . . 934.1 Formalism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94

viii

Chapter 5 High Energy Polarisation . . . . . . . . . . . . . . . . . . . . . . . 975.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 975.2 Degree of polarisation of blazar . . . . . . . . . . . . . . . . . . . . . . . . . 1005.3 Minimum Detectable Polarisation (MDP) - Detection sensitivity of telescopes 1045.4 Flaring . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1095.5 Indirect constraints on hadronic emission . . . . . . . . . . . . . . . . . . . . 1135.6 Discussion and conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115

Chapter 6 Radio Supernovae in the Great Survey Era . . . . . . . . . . . . 1176.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1186.2 Radio Properties of Supernovae . . . . . . . . . . . . . . . . . . . . . . . . . 120

6.2.1 Radio Core-Collapse Supernovae . . . . . . . . . . . . . . . . . . . . . 1206.2.2 Radio Type Ia Supernovae . . . . . . . . . . . . . . . . . . . . . . . . 122

6.3 Next-Generation Radio Telescopes: Expected Sensitivity . . . . . . . . . . . 1236.4 Radio Supernovae for SKA . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124

6.4.1 Core-Collapse Supernovae . . . . . . . . . . . . . . . . . . . . . . . . 1246.4.2 Type Ia Supernovae . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131

6.5 Radio Survey Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . 1326.6 Discussion and Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . 1346.7 Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136

Chapter 7 Conclusions and Future Work . . . . . . . . . . . . . . . . . . . . 1377.1 Discussion and Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . 1377.2 Future work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140

Appendix A . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142A.1 The Narrow Resonance Approximation . . . . . . . . . . . . . . . . . . . . . 142

A.1.1 Narrow Normal Resonances . . . . . . . . . . . . . . . . . . . . . . . 143A.1.2 Narrow Subhreshold Resonances . . . . . . . . . . . . . . . . . . . . . 144

Appendix B . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146B.1 The energy loss rates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148

ix

Chapter 1

Introduction

Nuclear and particle astrophysics comprises of applying standard principles of nuclear and

particle physics to astronomical systems and making predictions for observation. Also, ex-

treme environments in astronomical systems often serve as a probe for re-evaluating fun-

damental principles of physics. The research in this thesis comprises of both these aspects

of nuclear and particle astrophysics. One of the first steps in exploring new fundamental

physics with astrophysical probes, is to attain a systematic and comprehensive understanding

of the standard astrophysical processes and sources. Only after modeling these astrophysical

“foregrounds” can we confidently constrain models beyond the Standard Model of particle

physics and cosmology. With this broad vision in mind, I have worked on various projects

in the areas of primordial nucleosynthesis, the cosmic gamma-rays, high energy polarisation,

forecasting radio supernovae discovery, etc.

1.1 Primordial nucleosynthesis

As a part of the first topic, I have worked on the “Cosmological Lithium Problem”. The big

bang produces light elements namely H, He, Li and isotopes and isobars. And comparison of

theoretical predictions about the production of these elements with observations constrains

big bang cosmology (Wagoner et al., 1967a; Kolb & Turner, 1990; Cyburt, 2004a; Steigman,

2007, etc.). Big bang nucleosynthesis represents the competition between rate of nuclear,

weak and electromagnetic reactions, Γ and the expansion rate of the universe, or the Hubble

rate, H = a/a. Here a is the scale factor by which is unity at the present moment. The

1

reactions proceed to produce and destroy nuclei and achieve nuclear statistical equilibrium

between the various production and destruction reactions. This depends on the temperature

of the universe and also the baryon nb and photon densities nγ . In equilibrium, the rate of

interactions per particle exceed the expansion rate as

Γ = nσv ≫ H (1.1)

where σ is the interaction cross-section. Nucleosynthesis is quantified by through abun-

dances. And abundance of an element i is defined in terms of mole fraction Yi = ni/nb, with

respect to the baryon density of the universe at the time. The abundances are defined to be

the mole fractions, given relative to hydrogen, as

Yi

YH

=ni/nb

nH/nb

(1.2)

The dilution in the amount of these elements due to the expansion is precisely calculable

and a common factor for all nuclei. Yi is thus independent of this dilution factor. More

quantitatively the nuclear statistical equilibrium for a nucleus with abundance Y can be

quantified in terms of the abundances of projectiles, Yproj that produce it at a rate Γprod and

destruction rates, Γdest

dY

dt= nb(

prod

YprojΓprod −∑

dest

Y Γdest)

=>a

Yeq

dY

da= − Γ

H

((

Y

Yeq

)2

− 1

)

(1.3)

and the equilibrium abundance corresponding to Γ ≫ H is

Yeq =

YprojΓprod∑

Γdest

(1.4)

2

These abundances are functions of the temperature and also the baryon-to-photon ratio. Now

as the universe undergoes accelerated expansion, the expansion rate exceeds the production

and destruction reactions. As a result of this, different reactions fall out of equilibrium, and

the abundances set by those reactions are held fixed aside from the obvious dilution due to

expansion. This is falling out of equilibrium is known as “freeze-out” and the condition is

given as

Γ ≈ H (1.5)

Once, the expansion rate, H far exceeds reaction rate, Γ the abundances, Y are fixed. These

frozen out abundances depend only upon the baryon-to-photon density ratio represented by

η = nb/nγ ∝ Ωb/Ωγ. Here the Ω′s are ratio of matter-energy density of that component to

the critical density required to close the universe i.e. keep it from expanding forever.

Nucleosynthesis occurs at MeV energies and finishes within the first 3 minutes post the

big bang. It begins with protons and neutrons combining to produce deuterium. Weak inter-

actions interconvert neutrons and protons and when expansion exceeds the weak interaction

scale, the weak freezeout occurs setting the neutron to proton ratio(

np

)

∼ 1/6. The neutron

decays with a mean lifetime of ∼ 882 seconds. So a further decay of neutrons reduces this

ratio to 1/7, and this is when nucleosynthesis begins at the scale of the weak freezeout i.e.

at MeV temperatures. However, in addition to nucleons, there are about a billion times as

many photons in the radiation dominated universe at the time. This prevents neutrons and

protons from combining effectively. As a result, deuterium is photo-disintegrated as soon as

it is produced. This continues till about 100 seconds, at which point the universe cools to be-

low the binding energy of deuterium and “bottleneck” is released. Nucleosynthesis proceeds

to produce He and Li. The abundances of these elements after their respective freezeouts

are functions of the baryon-to-photon ratio alone. Therefore, observations of the elemental

abundances put constraints on this baryon-to-photon ratio other than parameters like the

3

gravitational constant, number of relativistic species, etc. that are fixed within the Standard

Model of nuclear and particle physics and cosmology. And through the baryon-to-photon

ratio or simply the baryon density, these abundances put constraints on big bang cosmology

(Kolb & Turner, 1990).

Further advances in the measuring the cosmic microwave background (CMB) have made

it a more precise baryometer than big bang nucleosynthesis (BBN) (Kneller et al., 2001;

Cyburt et al., 2003; Cuoco et al., 2004). Using the CMB value of Ωb, it possible to make

predictions for the elemental abundances of elements deuterium, tritium, helium and lithium

and isobars and isotopes. And these can be compared with observations. Thus, within the

framework of the Standard Model of particle physics and cosmology, there is no room /

freedom for adjusting the theoretical predictions of light elemental abundances.

Ideally, one would hope to observe these elements in pristine sites unpolluted by post

BBN- processing. However, in practice the observations are often affected by non-primordial

processes as stellar nucleosynthesis (For eg., Korn et al., 2006), cosmic rays and other as-

tronomical processes. Constraints from CMB are either weak as in the case of 4He or

unattainable due to the very nature of recombination. So the observations need to be per-

formed at the metal poor astronomical sites, minimally polluted by the heavier elements

or “metals” produced by aforementioned processes. 4He is observed in metal poor compact

dwarfs (Izotov et al., 1999). D is constrained from quasar absorption lines (O’Meara et al.,

2006; Fumagalli et al., 2011). 3He is observed in Galactic HII regions (Bania et al., 2002).

Observations of 7Li are performed at multiple sites as metal poor halo field stars (For e.g.

Spite & Spite, 1982; Ryan et al., 2000; Asplund et al., 2006) , globular cluster stars (Bonifa-

cio et al., 2002; Bonifacio, 2002; Gonzalez Hernandez et al., 2009) and more recently, in the

Small Magellanic Cloud (Howk et al., 2012). The 7Li observations in metal poor halo stars

show signs of a plateau called the “Spite Plateau” with a small scatter consistent with ob-

servational uncertainties. This suggests that the abundance is primordial. The observations

(Ryan et al., 2000) however, show an abundance 3-4 times less (Cyburt et al., 2008) than

4

predicted by standard BBN theory. This is known as the cosmological lithium problem.

One possible solution to the lithium problem is to look at nuclear reactions destroying

7Li or 7Be. 7Be is important as it is predominantly in this isobaric form that mass 7 nuclei

are formed in the early universe. Being β unstable it undergoes decay to produce the lithium

that is ultimately observed. The theoretical computation of the elemental abundances are

made by evolving the nuclear reactions both production and destruction, with the expansion

of the universe (Wagoner et al., 1967a; Smith et al., 1993). This in principle requires a

complete database of nuclear reactions whose properties are measured experimentally. In

practice, there are 12 important reactions including the neutron decay Smith et al. (1993).

Nuclear reactions that compete with these reactions that were either missed before or have

newer measurements will alter the abundances. Also, alternate nuclear channels, defined

by the quantum state of the reactants in these reactions, which were missed should also be

accounted for.

In Chakraborty et al. (2011), a systematic and complete study of these effects were made.

The 7Li and 7Be production reactions are very well measured leaving very little scope for

further exploration. So only destruction reactions need to be considered. It was found from

a semi-analytic estimate that only nuclear resonances were competitive enough with the 12

key reactions to alter the 7Li or 7Be abundances. This is based on the equilibrium abundance

in eq. 1.4). Thereby, a complete, systematic study of such resonances destroying 7Li and 7Be

down to observed abundances was made. The existence of such a resonance would represent

a nuclear solution to the cosmological 7Li problem that doesn’t require any new physics

beyond the Standard Model of particle physics and cosmology.

In doing so, data on resonances was used from resources such as NACRE 1, TUNL 2,

NNDC 3, etc. to construct resonant reaction rates to expand the nuclear reaction network

(Caughlan & Fowler, 1988; Angulo et al., 1999), in addition to the existing non-resonant

1http : //pntpm3.ulb.ac.be/Nacre/barre database.htm2http : //www.tunl.duke.edu/nucldata/3http : //www.nndc.bnl.gov/

5

contributions. Resonant energies and widths were parametrised where data was not available

and constrained the parameter space where 7Li acquires observed values. In doing so I

also, computed the amounts of heavier elements like 9Be are not overproduced by these

destruction reactions. Three candidate resonant levels for the initial states 7Be + d, 7Be + t

and 7Be+3He were found to have inelastic exit channels including p,3He and 4He which could

not be immediately dismissed based on the data on channel energies and widths available

then. However, owing to large Coulomb barriers, they all needed large channel radii (a > 10

fm) for solving the 7Li problem which makes this very difficult. More recently, 7Be + d

channel was ruled out explicitly by new data. This strongly suggests that systematically

accounting for standard nuclear physics cannot by itself solve the 7Li problem. As a result,

the 7Li problem may point to physics beyond the Standard Model of particle physics and

cosmology. These results along with work of others (Cyburt & Pospelov, 2012; Boyd et al.,

2010a) have prompted experiments (Charity et al., 2011; Kirsebom & Davids, 2011; O’Malley

et al., 2011) to look for these resonances. And these experiments have ruled out the best

candidate(s) as solutions to the lithium problem.

In absence, of this solution, it is a possibility that new physics beyond the Standard Model

has to be invoked in order to find a theoretical solution to the lithium problem. Several ways

of destroying exist from dark matter decay (Jedamzik, 2004; Jedamzik & Pospelov, 2009;

Ellis et al., 2005, etc.), variation of fundamental constants like quark masses (Dmitriev et al.,

2004), neutron lifetime, (Coc et al., 2007, etc.), etc., to formation of bound states (Cyburt

et al., 2006, etc.). Another class of solutions lies in stellar depletion of lithium in stars via

rotational mixing, turbulence, diffusion, etc (For e.g. Melendez et al., 2010b). Aside from

ways of destroying lithium, a reinterpretation of the observations due to an improved model

of stellar atmospheres may modify and reconcile the observed abundance with the theoretical

predictions (For e.g. Melendez et al., 2010a)

6

1.2 Cosmic gamma-rays

High energy radiation provides a natural, astrophysical probe to the some of the highest

energy phenomena and particles in the universe, all the way upto Planck scales. Thus,

cosmic gamma rays emitted by various astronomical sources both cosmological and local are

a powerful probe to the high energy universe. With the advances in gamma-ray astronomy

and the current state of the art with Fermi -LAT, it has become possible to investigate

individually interesting sources as well as populations of sources such as star forming galaxies

(Abdo et al., 2010f,a, etc.,), active galaxies or active galactic nuclei (Nolan et al., 2012a)

(AGN), compact objects like gamma-ray bursts (Fermi-LAT Collaboration, 2013), ordinary

black holes, etc. Now, in addition to known sources, gamma-rays are probes to the unknown

and the exotic. Just as BBN, cosmic gamma-rays too are an indirect probe of particle dark

matter (Ando et al., 2007; Regis & Ullio, 2008; Ackermann et al., 2012b). Dark matter decays

and annihilations produce gamma-rays at various energies and studying their properties

gives clues about dark matter models. However, it is crucial as in the case of BBN to

eliminate all the Standard Model foregrounds that could mimic or distort the dark matter

signature. Or alternately, we must model all the standard astrophysical sources that can

explain the data before invoking physics beyond Standard Model as would be the case for

dark matter. Therefore, even in the study of the dark matter, it is important to study

the known sources mentioned before. Various properties of gamma-ray sources such as

source spectra (Stecker & Venters, 2011; Lacki et al., 2012; Fields et al., 2010; Ackermann

et al., 2012c), flux distribution and statistics (Abdo et al., 2010l; Malyshev & Hogg, 2011),

anisotropies (Ackermann et al., 2012a; Hensley et al., 2010), etc. have been studied in order

to firstly, understand the distinctive signatures of the various astrophysical sources producing

gamma-rays. This is achieved by studying these properties of the resolved sources. Secondly,

EGRET (Sreekumar et al., 1998) and now Fermi -LAT (Abdo et al., 2010j) has helped

confirm the existence of a diffuse, extra-Galactic gamma-ray background (EGB). Efforts are

on to try to identify the origin of this diffuse EGB. Blazars or AGNs with their jets pointed

7

towards the observer, have been identified as the being the dominant source class amongst

the resolved sources. It is therefore natural to determine their contribution to the unresolved

sources or diffuse background. It is clear from the results of various groups including ours

(For eg Chakraborty & Fields, 2012; Stecker & Venters, 2011; Malyshev & Hogg, 2011)

that the diffuse background is multi-component. It is unlikely to be dominated by AGNs

due to guaranteed contribution from other sources mainly star forming galaxies Pavlidou &

Fields (2002); Fields et al. (2010). Despite knowing this, there is still debate on the exact

contribution of the star forming galaxies and AGNs, the two guaranteed sources of the EGB.

It is important to sharpen the constraints on their relative contributions.

As individuals objects, star-forming galaxies are laboratories of cosmic ray physics with

X-ray and gamma-ray telescopes being the apparatus. High energy radiation is produced in

sites of supernova remnants (SNR) due to cosmic ray interactions. Therefore, the gamma-

rays produced for instance are cosmic gamma rays. Multiwavelength observations provide

evidence of cosmic ray acceleration in sites of SNR upto energies of ∼ 1015 eV (Uchiyama

et al., 2007; Helder et al., 2009, etc.,). The cosmic ray flux φCR is the proportional to the

supernova rate. The supernova rate in turn, is proportional to the massive star-formation

rate, which ties it to the gamma-ray luminosity in sites of SNR in star-forming galaxies.

There exist direct connections between cosmic gamma rays and star-formation. This con-

nection is manifested in the context of the gamma-ray luminosities of individual galaxies

as well as the diffuse extra-Galactic gamma ray background (EGB).Star-forming galaxies

produce gamma-rays by cosmic ray interactions with the interstellar medium (ISM). And

therefore, the gamma-ray emissivity, qγ is given in terms of the cosmic ray flux φCR and the

ISM density nISM as,

qγ ∼ φCRσintnISM (1.6)

Gamma ray luminosities of individual star-forming galaxies are correlated to the galactic

8

star-formation rate and the cosmic star-forming background to the cosmic star-formation

rate. The exact nature of the correlation depends on the emission mechanism, hadronic or

leptonic and their efficiencies.

Star-forming galaxies from the Local Group have been detected at GeV energies by Fermi

-LAT (Abdo et al., 2010g). In addition to the Milky Way this includes normal galaxies

M 31(Abdo et al., 2010g), Large Magellanic Cloud (LMC) (Abdo et al., 2010h), Small

Magellanic Cloud (SMC) (Abdo et al., 2010c) and starbursts M 82 and NGC 253 (Abdo

et al., 2010b). M33 is very likely to be detected by Fermi -LAT in the upcoming years

(Abdo et al., 2010g). It is believed that the dominant emission mechanism for Milky Way

type galaxies GeV gamma-ray production is hadronic or due to interactions of cosmic ray

protons with interstellar gas. (neutral, ionised and molecular hydrogen) to produce neutral

pion that decay in flight into gamma rays. The gamma-ray emission can be viewed in terms

of collisions between projectiles i.e. the cosmic ray protons with the targets i.e. the ISM

protons.

pCR + pISM → p+ p+ π0 π0 → γγ (1.7)

However, in some cases as in the Magellanic Clouds, this is not clear as the gas and the

gamma rays show very little spatial correlation. The next most important process of pro-

ducing gamma-rays is inverse-Compton scattering of cosmic ray electrons off the interstellar

radiation field. Here the projectiles are the cosmic ray electrons and the targets are the ISRF

photons.

e−CR + γISRF → e− + γHE (1.8)

In order to determine the exact contribution of star forming galaxies, it is important

to study the individual galaxies in detail as only the handful mentioned above have been

detected in gamma-rays Abdo et al. (2010g) in contrast the vast majority of AGNs. There-

9

fore, while for AGNs a statistical understanding of the gamma-ray emission suffices, for star

forming galaxies knowing details of the emission mechanisms in these handful are useful

to extrapolate to get their cosmological emission. Emission mechanisms of the individual

galaxies as well as the number density of these galaxies differ depending on whether they are

a starburst that has heightened periods of star formation activity or not. The ones which

dont have this starburst activity are called normal galaxies. Pavlidou & Fields (2002); Fields

et al. (2010), computed a model for the hadronic emission due to neutral pion decay from

normal star forming galaxies. This involves computation of the projectile or cosmic ray flux

in terms of the star formation rate based on the aforementioned correlation and normalised

to the Milky Way. The target density is the interstellar hydrogen density in various forms

i.e. neutral, molecular and ionised. The relevant quantity here is the surface gas density that

in turn is related to the surface density of star formation Kennicutt (1998). This implies

that the gamma-ray luminosity Lγ is a non-linear function of the star formation rate ψ. This

is a critical result from Pavlidou & Fields (2002); Fields et al. (2010) that sets the overall

scaling relation between the gamma-ray luminosity and the star formation rate. The shape

of course, is very well defined. A pion decaying at rest would have a delta function peak

at the rest mass of 67.5 MeV. However, the pion decays in flight. Furthermore, Galactic

rotation produces dispersion, and the peak becomes a bump. For external galaxies at cos-

mic distances, there is a redshifting as well. Using Hα as a tracer for these normal star

forming galaxies, the distribution function is constructed. The intensity is then computed

as the line of sight integral of product of the distribution as well as the luminosity of a single

galaxy. Fields et al. (2010) find that the cosmological contribution of the normal galaxies

due to neutral pion decay is significant and could in fact dominate the extra-Galactic diffuse

background upto around 10 GeV. The LAT range of course, extends to around 300 GeV.

Given that the inverse-Compton contribution from normal galaxies is also guaranteed

and the next most important, this contribution was computed in Chakraborty & Fields

(2012). Also, since, the inverse Compton spectral shape is flatter at higher energies it

10

could apriori be significant at energies above 10 GeV. This provides strong motivation to

perform this computation. This leptonic contribution has cosmic ray electrons as projectiles

instead of protons. Once again as for protons, the cosmic ray electron flux scales as the

star formation rate, being accelerated in the same SNR as protons. Their efficiency could

be different though. Electrons propagate differently from the protons. Radiative losses are

significant for electrons, unlike protons that are heavier. Also, while electrons diffuse they

do not escape the galaxy whereas proton losses in the normal galaxies are dominated by

escape. The radiative energy lost by electrons is equal to the energy output in photons at

various energies depending on loss mechanism. This is known as calorimetry. For instance,

synchrotron losses produce photons from radio all the way to X-rays. Inverse Compton losses

produce some X-rays and certainly gamma-rays. And thus the gamma-ray IC emission is

basically the fraction of the total energy lost in IC. As a result the gamma-ray luminosity

depends not the the background seed photon or ISRF density, but fraction of IC. Once

again from the luminosity and the Hα distribution the intensity of cosmological emission is

computed.

1.3 High energy polarisation

In section 1.2, gamma-ray properties of star forming galaxies and AGNs discussed include

the source intensity, spectra, flux distribution, etc. A property with tremendous scientific

potential in gamma-ray and even X-ray astronomy that is being explored more recently

(For e.g. McNamara et al., 2009; Krawczynski, 2012; Weisskopf et al., 2010) is polarisation.

Measurement of polarisation at X-ray and even soft gamma-rays is very challenging. But

several experiments in various stages of operation, design and commission are about to change

this and open a new window of opportunities in investigating extreme objects like AGNs,

GRBs, black holes in general, neutron stars, etc (McNamara et al., 2009; Krawczynski,

2012; Gotz et al., 2009). Some of the above sources are either established as or candidates

11

for particle accelerators all the way upto nearly the ankle of the cosmic ray spectrum (∼

1019eV). HIgh energy polarisation will probe of emission mechanisms that in turn will probe

acceleration mechanisms and thereby test the particle accelerators. Magnetic fields can

lead to polarised emission. And therefore, polarisation measurements will help understand

magnetic field structure surrounding the astronomical sources.

Lorentz invariance is one of the treasured symmetries in the Standard Model of particle

physics and cosmology. But certain particle physics and gravity models have Lorentz invari-

ance violation built into them. Gamma rays can be used to probe these models in many ways

(Coleman & Glashow, 1999; Ellis et al., 2000) One of the ways is to measure vacuum bire-

fringence, or modification of the dispersion relation of light due to Lorentz violating terms

in the underlying Lagrangian. Thus, putting constraints on these terms puts constraints

on Lorentz violation. This modified dispersion relation leads to a rotation of polarisation

during propagation of the photons. INTEGRAL / IBIS observations of the polarisation from

the prompt emission from the GRB041219A has put a strong limit on the cubic correction

term to the standard dispersion relation (Laurent et al., 2011)limiting the LIV parameter to

< 1.110×10−14, an improvement of 4 orders of magnitude over the previous limit. This limit

is inversely proportional to the distance, which suggests that polarisation measurements of

higher redshift GRBs can in principle improve the limits further.

In particular, polarised X-rays and gamma-rays from AGNs have the potential to reveal a

lot of different aspects on AGN physics. AGNs are some of the most luminous yet mysterious

astronomical objects in the universe. Their particle and radiative emissions are powered by a

central supermassive black hole accreting matter. Part of the gravitational energy associated

with accretion is converted into energy of particles such as cosmic rays and neutrinos and

high energy radiation like X-rays and gamma-rays. Thus, these particles and radiation are

messengers of the extreme astrophysical conditions in the core of active galaxies. Blazars are

distant AGNs where the observer’s line of sight is along the jet axis, i.e. the observer looks

down into the jets. Various properties of the radiation from blazars like the overall inten-

12

sity, spectrum, variability have been studied with multiwavelength observations. However,

polarisation particularly at high energies has received much less attention. However, with

numerous X-ray and soft gamma-ray polarimeters at various stages of planning, design and

operation and studies of optical / FIR polarisation properties of blazars underway, this is a

perfect time to calculate high energy polarisation from blazars. High energy polarisation is

a key ingredient in the multimessenger, multiwavelength understanding of blazars.

1.4 Radio Supernovae

Yet another piece in this gamma-ray - star formation connection are core collapse SNe.

Core collapse SNe are result of death of massive stars. And hence, the core collapse SNe

rate traces the massive star formation rate. In fact, they are in direct proportion to each

other. Thus building up a statistically significant sample of core-collapse SNe constitutes

an independent probe of the star formation rate. Core collapse SNe have been observed

to emit radio. Over 50 of them have been discovered with the current generation of radio

telescopes. However, with future generations of radio surveys that represent more than an

order of magnitude improvement in sensitivity, it is possible to discover many more core

collapse SNe. This discovery potential is explored in Lien et al. (2011). Radio has the

capability to go deeper than optical in discovering supernovae. The large optical telescopes

such as the Large Synoptic Survey Telescope (LSST) will detect a plethora of core-collapse

supernovae out to redshift 1. However, it is estimated by Mannucci et al. (2007) that

∼ 60% of core collapse supernovae maybe missed in optical surveys due to increased dust

obscuration at high redshift. In our work in Lien et al. (2011) , it was shown that future

generations of radio telescopes in survey mode such as SKA and its precursors will detect

a large fraction of core-collapse SNe as predicted from the cosmic massive star-formation

rate in Horiuchi et al. (2009a) out to redshift 5. Also, future radio telescope arrays could

be trained with simultaneous or triggered with preceding optical detections at low redshift.

13

And using this knowledge about radio supernovae at low redshift, they could be used for

independent discoveries at high redshift. Also, there is a potential shift in the detection

strategies from mainly targeted PI driven searches to untargeted, synoptic surveys capable

of detecting radio supernovae automatically.

The radio emission comes from the interaction of the wind with the circumstellar medium.

So other simply building a sample, naturally, the radio emission is a probe of the models of

the circumstellar medium and its various parts Chevalier (1982a,b).

About 25% of core-collapse SNe are Type 1bc Li et al. (2011a), which are related to

long GRBs. According to our estimate with very conservative assumptions about fraction

of radio emitting Ibc SNe, we expect SKA to make ∼ 130 unbiased, untargeted detections,

with ∼ 20 connected with GRBs. Aside from these, exotic transients whose properties do

not match those of known transient sources are yet another discovery pool.

A synoptic survey in radio wavelengths will be crucial in many fields of astrophysics,

particularly for RSNe, but also for other transients. It will bring the first complete and

unbiased RSN sample and systematically explore exotic radio transients. SKA will be ca-

pable of performing such an untargeted, possibly automated survey with its unprecedented

sensitivity. Our knowledge of supernovae will thus be firmly extended into the radio and to

high redshifts.

14

Chapter 2

Resonant Destruction of LithiumDuring Big Bang Nucleosynthesis

Abstract

1 We explore a nuclear physics resolution to the discrepancy between the predicted standard

big-bang nucleosynthesis (BBN) abundance of 7Li and its observational determination in

metal-poor stars. The theoretical 7Li abundance is 3 - 4 times greater than the observational

values, assuming the baryon-to-photon ratio, ηwmap, determined by WMAP. The 7Li problem

could be resolved within the standard BBN picture if additional destruction of A = 7 isotopes

occurs due to new nuclear reaction channels or upward corrections to existing channels. This

could be achieved via missed resonant nuclear reactions, which is the possibility we consider

here. We find some potential candidate resonances which can solve the lithium problem and

specify their required resonant energies and widths. For example, a 1− or 2− excited state of

10C sitting at approximately 15.0 MeV above its ground state with an effective width of order

10 keV could resolve the 7Li problem; the existence of this excited state needs experimental

verification. Other examples using known states include 7Be + t → 10B(18.80 MeV), and

7Be+d→ 9B(16.71 MeV). For all of these states, a large channel radius (a > 10 fm) is needed

to give sufficiently large widths. Experimental determination of these reaction strengths is

needed to rule out or confirm these nuclear physics solutions to the lithium problem.

1This chapter is previously published in The Physical Review D as Chakraborty, N., Fields, B. D., &Olive, K. A. 2011, Phys. Rev. D, vol. 83 pp. 63006. This chapter matches the published version aside fromsuperficial modifications to references.

15

2.1 Introduction

Primordial nucleosynthesis continues to stand as our earliest probe of the universe based

on Standard Model physics. Accurate estimates of the primordial abundances of the light

elements D, 4He and 7Li within standard Big Bang Nucleosynthesis (BBN) Cyburt et al.

(2001); Coc et al. (2004a); Cyburt et al. (2003); Cyburt (2004b); Cyburt et al. (2008) are

crucial for making comparisons with observational determinations and ultimately testing the

theory. Primordial abundances are also a probe of the early universe physics Cyburt et al.

(2005). Currently, the theoretical estimates of D and 4He match the observational values

within theoretical and observational uncertainties Cyburt et al. (2003, 2008) at the baryon-

to-photon ratio determined by the 7-year WMAP data, ηwmap = 6.19±0.15×10−10 Komatsu

et al. (2011). In contrast, the theoretical primordial abundance of 7Li does not match the

observations.

At ηwmap, the predicted BBN abundance of 7Li is2 Cyburt et al. (2008)

(

7Li

H

)

BBN

=(

5.12+0.71−0.62

)

× 10−10. (2.1)

The observed 7Li abundance is derived from observations of low-metallicity halo dwarf stars

which show a plateauSpite & Spite (1982) in (elemental) lithium versus metallicity, with a

small scatter consistent with observational uncertainties. An analysis Ryan et al. (2000) of

field halo stars gives a plateau abundance of

(

Li

H

)

halo⋆

= (1.23+0.34−0.16) × 10−10. (2.2)

However, the lithium abundance in several globular clusters tends to be somewhat higher Boni-

facio (2002); Gonzalez Hernandez et al. (2009), and a recent result found in Gonzalez

Hernandez et al. (2009) gave 7Li/H = (2.34 ± 0.05) × 10−10. Thus the theoretically es-

2Note that the 7Li abundance reported here differ slightly from that given in Cyburt et al. (2008),primarily due to the small shift in η as reported in Komatsu et al. (2011).

16

timated abundance of the isobar with mass 7 (7Be+7Li) is more than the observationally

determined value by a factor of 2.2 - 4.2 Cyburt et al. (2008), at ηwmap. Relative to the

theoretical and observational uncertainties, this represents a deviation of 4.5-5.5 σ.

This significant discrepancy constitutes the “lithium problem” which could point to lim-

itations in either the observations, our theoretical understanding of nucleosynthesis, or the

post-BBN processing of lithium.

On the theoretical front, strategies which have emerged to approach the lithium problem

broadly either address astrophysics or microphysics. On the astrophysical side, one might

attempt to improve our understanding of lithium depletion mechanisms operative in stellar

models Korn et al. (2006). This remains an important goal but is not our focus here.

The microphysical solutions to the lithium problem all in some way change the nuclear

reactions for lithium production in order to reduce the primordial (or pre-Galactic) lithium

abundance to observed levels. Some of these work within the Standard Model, focussing on

nuclear physics, in particular the nuclear reactions involved in lithium production. One ap-

proach is to attempt to utilize the experimental uncertainties in the rates Coc et al. (2004a);

Angulo et al. (2005); Cyburt et al. (2004); Boyd et al. (2010b). A second, related approach

is the inclusion of new effects in the nuclear reaction database such as poorly understood

resonance effects Cyburt & Pospelov (2012). Finally, it may happen that effects beyond

the Standard Model are responsible for the observed lithium abundance. For example, the

primordial lithium abundance can be reduced by cosmological variation of the fine structure

constant associated with a variation in the deuterium binding energy Coc et al. (2007), or

by the post-BBN destruction of lithium through the late decays of a massive particle in the

early universe Cyburt et al. (2010).

In this paper, we remain within the Standard Model, examining the possible role of

resonant reactions which may have been up to now neglected. The requisite reduction in the

7Li abundance can be achieved by either an enhancement in the rate of destruction of 7Li

or its mirror nucleus 7Be. This approach is more promising than the alternative of reducing

17

the production of 7Be and 7Li where the reactions are better understood experimentally

and theoretically Cyburt (2004a); Cyburt & Davids (2008); Ando et al. (2006), whereas the

experimental and especially the theoretical situation for A = 8 − 11 has made large strides

but still allows for surprises at the levels of interest to us Pieper et al. (2002).

The use of resonant channels is an approach that has paid off in the past in the context of

stellar nucleosynthesis. Fred Hoyle famously predicted a resonant energy level at 7.68 MeV

in the 12C compound nucleus which enhances the 8Be + α→ 12C reaction cross-section and

allows the triple alpha reaction to proceed at relatively low densities Hoyle (1954). Recently,

it was shown that there are promising resonant destruction mechanisms which can achieve

the desired reduction of the total A = 7 isotopic abundance Cyburt & Pospelov (2012). This

paper points to a resonant energy level at (E, Jπ) = (16.71 MeV, 5/2+) in the 9B compound

nucleus which can increase the rate of the 7Be(d,p)αα and/or 7Be(d, γ)9B and thereby reduce

the 7Be abundance. Here, we take a more general approach and systematically search for all

possible compound nuclei Tilley et al. (2004); Ajzenberg-Selove (1990) and potential resonant

channels which may result in the destruction of 7Be and/or 7Li.

Because of the large discrepancy between the observed and BBN abundance of 7Li, any

nuclear solution to the lithium problem will require a significant modification to the ex-

isting rates. As we discuss in the semi-analytical estimate in section 2.2, any new rate or

modification to an existing one, must be 2 - 3 times greater than the current dominant

destruction channels namely, 7Li(p, α)α for 7Li and 7Be(n, p)7Li for 7Be. As discussed in

Boyd et al. (2010b) and as we show semi-analytically in § 2.2, this is difficult to achieve

with non-resonant reactions. Hence, we will concentrate on possible resonant reactions as

potential solutions to the lithium problem. As we will show, there are interesting candidate

resonant channels which may resolve the 7Li problem. For example, there is a possibility of

destroying 7Be through a 1− or 2− 10C excited state at approximately 15.0 MeV. The energy

range between 6.5 and 16.5 MeV is currently very poorly mapped out and a state near the

entrance energy for 7Be + 3He could provide a solution if the effective width is of order 10

18

keV. We will also see that these reactions all require fortuitously favorable nuclear param-

eters, in the form of large channel radii, as also found by Cyburt and Pospelov Cyburt &

Pospelov (2012) in the case of 7Be+d. Even so, in the face of the more radical alternative of

new fundamental particle physics, these more conventional solutions to the lithium problem

beckon for experimental testing.

The paper is organized as follows: First, we lay down the required range of properties of

any resonance to solve the lithium problem by means of a semi-analytic estimate inspired

by Mukhanov (2004); Esmailzadeh et al. (1991) in § 2.2. Then, in § 2.3, we list experimen-

tally identified resonances from the databases: TUNL 3 and NNDC 4, involving either the

destruction of 7Be or 7Li. Finally, the solution space of resonant properties, wherein the

lithium problem is partially or completely solved, is mapped for the most promising initial

states involving either 7Li or 7Be, by including these rates in a numerical estimation of the

7Li abundance. This exercise will delineate the effectiveness of experimentally studied or

identified resonances as well as requirements of possible missed resonant energy levels in

compound nuclei formed by these initial states. This is described in § 2.4. We note that

in our analysis, the narrow resonance approximation is assumed which may not hold true

in certain regions of this solution space. Our key results are pared down to a few resonant

reactions described in § 2.5. A summary and conclusions are given in § 2.6.

2.2 Semi-analytical estimate of important reaction

rates

Before we embark on a systemic survey of possible resonant enhancements of the destruction

of A = 7 isotopes, it will be useful to estimate the degree to which the destruction rates must

change in order to have an impact on the final 7Li abundance. The net rate of production

3http : //www.tunl.duke.edu/nucldata/4http : //www.nndc.bnl.gov/

19

of a nuclide i is given by the difference between the production from nuclides k and l and

the destruction rates via nuclide j, i.e. for the reaction i+ j → k + l. This is expressed

quantitatively by the rate equation Wagoner et al. (1967b) for abundance change

dni

dt= −3Hni +

jkl

nknl〈σv〉kl − ninj〈σv〉ij, (2.3)

where ni is the number density of nuclide i,H is the Hubble parameter,∑

ij ninj〈σv〉ij are the

sum of contributions from all the forward reactions destroying nuclide i and∑

kl nknl〈σv〉kl

are the reverse reactions producing it. 〈σv〉 is the thermally averaged cross-section of the

reaction. The dilution of the density of these nuclides due to the expansion of the universe

can be removed by re-expressing eq. (2.3) in terms of number densities relative to the baryon

density Yi ≡ ni/nb, as,

dYi

dt= nb

jkl

YkYl〈σv〉kl − YiYj〈σv〉ij . (2.4)

Using this general form, the net rate of 7Be production can be approximated in terms of the

thermally averaged cross-sections of its most important production and destruction channels

as

dY7Be

dt= nb (〈σv〉3HeαY3HeYα − 〈σv〉7BenY7BeYn) . (2.5)

Here, the reverse reaction rates of these production and destruction channels are neglected,

as they are much smaller than the forward rates at the lithium synthesis temperature. A

similar equation can be written down for 7Li. When quasi-static equilibrium is reached, the

destruction and production rates are equal. In this case, approximate values for new rates,

which can effectively destroy either isobar, can be obtained analytically.

At temperatures T ≈ 0.04 MeV, both 7Li and 7Be are in equilibrium Mukhanov (2004)

20

which gives,

〈σv〉3HeαY3HeYα = 〈σv〉7BenY7BeYn . (2.6)

Consider a new, inelastic 7Be destruction channel 7Be+X → Y +Z, involving projectile X.

This reaction will add to the right hand side of eq. (2.6) and shift the equilibrium abundance

of 7Be to a new value as follows,

Y new7Be ≈ 〈σv〉3HeαYαY3He

〈σv〉7BenYn + 〈σv〉7BeXYX≈ 1

1 +〈σv〉7BeXYX

〈σv〉7BenYn

Y old7Be . (2.7)

If the new reaction is to be important in solving the lithium problem, it must reduce the

7Be abundance by a factor of Y new7Be

/Y old7Be

∼ 3 − 4 . This in turn demands via eq. (2.7) that

〈σv〉7BeXYX/〈σv〉7BenYn ∼ 2 − 3, i.e., the rate for the new reaction exceeds that of the usual

n− p interconversion reaction rate. A similar estimate can be made for 7Li.

This reasoning would exclude non-resonant rates as they would be required to have

unphysically large astrophysical S-factors in the range of order 105−109 keV - barn depending

on the channel. Thus we would expect that only resonant reactions can produce the requisite

high rates. Possible resonant reactions are listed in the next section, whose key properties of

resonance strength, Γeff and energy, Eres, lie in appropriate ranges capable of achieving the

required destruction of mass 7.

Finally, we turn to 7Li destruction reactions, 7Li+X → Y +Z. Recall that at the WMAP

value of η, mass 7 is made predominantly as 7Be, with direct 7Li production about an order

of magnitude smaller. This suggests that enhancing direct 7Li destruction will only modestly

affect the final mass-7 abundance; we will see that this expectation is largely correct.5

With these pointers, the list in the next section is reduced and numerical analysis of the

5A subtle point is that normally, the mass-7 abundance is most sensitive to rate 7Be(n, p)7Li Smith et al.(1993). Of course, this reaction leaves the mass-7 abundance unchanged, but the lower Coulomb barrier for7Li leaves it vulnerable to the 7Li(p, α)4He reaction, which is extremely effective in removing 7Li. Thus, fora new, resonant 7Li destruction reaction to be important, it must successfully compete with the very large7Li(p, α)4He rate, and even then the mass-7 destruction “bottleneck” remains the 7Be(n, p)7Li rate thatlimits 7Li appearance. Thus we would not expect direct 7Li destruction to be effective. We will examine 7Lidestruction below, and confirm these expectations.

21

remaining promising rates is done.

2.3 Systematic Search for Resonances

In this section we describe a systematic search for nuclear resonances which could affect

primordial lithium production. We first begin with general considerations, then catalog the

candidate resonances. We briefly review the basic physics of resonant reactions to establish

notation and highlight the key physical ingredients.

2.3.1 General Considerations

Energetically, the net process 7Be + A → B +D must have Q+ Einit ≥ 0, where the initial

kinetic energy Einit ≃ T <∼ 40 keV is small at the epoch of A = 7 formation. Thus we in

practice require exothermic reactions, Q > 0. Moreover, inelastic reactions with large Q will

yield final state particles with large kinetic energies. Such final states thus have larger phase

space than those with small Q and in that sense should be favored.

Consider now a process 7Be + X → C∗ → Y + Z which destroys 7Be via a resonant

compound state; a similar expression can be written for 7Li destruction. In the entrance

channel 7Be + X → C∗ the energy released in producing the compound state is QC =

∆(7Be) + ∆(X) − ∆(Cg.s.), where ∆(A) = m− Amu is the mass defect. If an excited state

C∗ in the compound nucleus lies at energy Eex, then the difference

Eres ≡ Eex −QC (2.8)

determines the effectiveness of the resonance. We can expect resonant production of C∗

if Eres <∼ T . In an ordinary (“superthreshold”) resonance we then have Eres > 0, while a

subthreshold resonance has Eres < 0.

Once formed, the excited C∗ level can decay via some set of channels. The cross section

22

for 7Be +X → C∗ → Y + Z is given by the Breit-Wigner expression

σ(E) =πω

2µE

ΓinitΓfin

(E −Eres)2 + (Γtot/2)2(2.9)

where E is the center-of-mass kinetic energy in the initial state, µ is the reduced mass and

ω =2JC∗ + 1

(2JX + 1)(2J7 + 1)(2.10)

is a statistical factor accounting for angular momentum. The width of the initial state

(entrance channel) is Γinit, and the width of the final state (exit channel) is Γfin.

One decay channel which must always be available is the entrance channel itself. Obvi-

ously such an elastic reaction is useless from our point of view. Rather, we are interested

in inelastic reactions in which the initial 7Be (or 7Li) is transformed to something else. In

some cases, an inelastic strong decay is possible where the final state particles Y + Z are

both nuclei. Note that it is possible to produce a final-state nucleus in an excited state, e.g.,

C∗ → Y ∗+Z, in which case the energy release Q′C is offset by the Y ∗ excitation energy. This

possibility increases the chances of finding energetically allowable final states. Indeed, such

a possibility has been suggested in connection with the 7Be + d → 9B∗ → 8Be

∗+ p process

Cyburt & Pospelov (2012).

Regardless of the availability of a strong inelastic channel, an electromagnetic transition

C∗ → C(∗)+γ to a lower level is always possible. However, these often have small widths and

thus a small branching ratio Γfin/Γtot. Thus for electromagnetic decays to be important, a

strong inelastic decay must not be available, and the rest of the reaction cross section needs

to be large to compensate the small branching; as seen in eq. (2.9), this implies that Γinit be

large.

Note that in all charged-particle reactions, the Coulomb barrier is crucially important

and is implicitly encoded via the usual exponential Gamow factor in the reaction widths of

both initial and final charged-particle states. However, if the reaction has a high Q, the final

23

state kinetic energy will be large and thus there will not be significant final-state Coulomb

supression; this again favors final states with large Q. In addition, if the entrance or exit

channel has orbital angular momentum L > 0, there is additional exponential suppression,

so that L > 0 states are disfavored for our purposes.

With these requirements in mind, we will systematically search for resonant reactions

which could ameliorate or solve the lithium problem. We begin by identifying possible

processes which are

1. new resonances not yet included in the BBN code;

2. 2-body to 2-body processes, since 3-body rates are generally very small in BBN due

to phase space suppression as well as the relatively low particle densities and short

timescales;

3. experimentally allowed – in practice this means we seek unidentified states in poorly

studied regimes;

4. narrow resonances having Γtot <∼ T , which is around Γtot < 40 keV but we will also

consider somewhat larger widths to be conservatively generous.

5. relatively low-lying resonances with Eres <∼ few × T ∼ 100 − 300 keV, which are

thermally accessible; here again we err on the side of a generous range.

Once we have identified all possible candidate resonances, we will then assess their viability

as solutions to the lithium problem based on available nuclear data.

2.3.2 List of Candidate Resonances

As described above, we will explore the resonant destruction channels of both 7Li and 7Be.

Some of the potential resonances which might be able to reduce the mass 7 abundance to

the observed value were recently considered in Cyburt & Pospelov (2012). This analysis

eliminates several candidate resonances, leaving as genuine solutions only the resonance

24

related to the 7Be(d, γ)9B and 7Be(d, p)αα reactions and associated with the 16.71 MeV

level in the 9B compound nucleus. Here, we make an exhaustive list of possible promising

resonances that may be important to either 7Be or 7Li destruction channels. In order to

do so systematically and account for all possible resonances that may be of importance, we

study the energy levels in all possible compound nuclei that may be formed in destroying

7Be or 7Li, making extensive use of databases at TUNL and the NNDC 6.

The available 2-body destruction channels 7A+X may be classified byX = n, p, d, t, 3He, α,

and γ. Consequently, the compound nuclei that can be formed starting from mass 7 have

mass numbers ranging from A = 8 to A = 11, and the ones of particular interest are

8Li, 8Be, 8B, 9Be, 9B, 10Be, 10B, 10C, 11B and 11C. All relevant, resonant energy levels in these

compound nuclei that may provide paths for reduction of mass-7 abundance are listed in

Tables 2.1 – 2.10.

There are quantum mechanical and kinematic restrictions to our selection of candidates.

The candidate resonant reactions must obey selection rules. The partial widths for a channel,

which may be viewed as probability currents of emission of the particle in that channel

through the nuclear surface, are given as

ΓL(E) = 2 ka PL(E, a)γ2(a) (2.11)

where a is the channel radius and E is the projectile energy. Here k is the wavenumber of the

colliding particles in the centre-of-mass frame and γ2 is the reduced width, which depends on

the overlap between the wavefunctions inside and outside the nuclear surface, beyond which

the nuclear forces are unimportant. The reduced width, γ2 is independent of energy and has

a statistical upper limit called the Wigner limit given by Teichmann & Wigner (1952)

γ2 ≤ 3~2

2µa2, (2.12)

6http : //www.tunl.duke.edu/nucldata/,http : //www.nndc.bnl.gov/

25

The pre-factor of 32

is under the assumption that the nucleus is uniform and can change

to within a factor of order unity if this assumption changes. The Wigner limit depends

sensitively on the channel radius and thus varies with the nuclei involved. For the nuclei of

our interest, typical values of γ2 range from a few hundred keVs to a few MeVs.

In eq. (2.11), PL(E, a) is the Coulomb penetration probability for angular momentum L

and is a strong and somewhat complicated function of E and a. Thus, while the Wigner

limit sets a theoretical limit on the reduced width, the upper limit on the full width, ΓL(E),

depends on the values of PL(E, a) and is sensitive to the details of the resonant channel

being considered. In light of this complexity, our strategy is as follows. We evaluate the

ΓL(E) needed to make a substantial impact on the lithium problem. Then for the cases of

highest interest, we will compare our results with the theoretical limit set by the Coulomb

suppressed Wigner limit for those specific cases.

We also limit our consideration to two body initial states, with resonance energies Eres ≤

650 keV. The high resonance energy limit ensures that all possible resonances which may

influence the final 7Li abundance are taken into account, though many of the channels with

such high resonance energies will inevitably be eliminated. Excited final states have also been

considered in making this list. Different excited states of final state products are marked as

separate entries in the table, since each one has its own spin and therefore a different angular

momentum barrier. And thereby the significance of each excited state in destroying mass

7 is varied. Also, we usually eliminate the reactions with a negative Q-value except for the

7Li(d, p)8Li, 7Be(d, 3He)6Li, 7Be(d, p)8Be∗

(16.922 MeV) and 7Li(3He, p)9Be∗

(11.283 MeV)

as they are only marginally endothermic.

For a number of the reactions listed in these tables, 1-10, the total spin of the initial

state reactants is equal to that of the compound nucleus, which is equal to the total spin of

the products, with L = 0. However, for many reactions, angular momentum is required in

the initial and/or final state, which decreases the penetration probability and thereby the

width for that particular channel. In fact for some of these reactions, parity conservation

26

demands higher angular momentum which worsens this effect. However, we do not reject any

channel based on the angular momentum suppression of its width in these tables. Later we

will shortlist those resonant channels which are not very suppressed and indeed potentially

effective in destroying mass 7.

The reactions of interest are listed in increasing order of the mass of the compound nuclei

formed. The particular resonant energy levels of interest Eex and their spins are listed in the

table. In general, different initial states involving 7Li and 7Be can form these energy levels

and so all these relevant initial states are listed. For each one, the various final product

states for an inelastic reaction are enumerated. Again, each of the final state products can

also be formed in an excited state. These excited states must have lower energy than the

initial state energies for the reaction to be exothermic. In addition spin and parity must

be conserved. Enforcing these, the minimum final state angular momenta Lfin are evaluated

from the spin of the resonant energy level and are listed in the tables. The total widths of

the energy levels are listed whenever available. The partial widths of the different channels

including the elastic one, out of each energy level are also listed.

We adopt the narrow resonance approximation to evaluate the effect of these resonances

and either retain or dismiss them as potential solutions to the lithium problem. Some of

the partial widths or limits on them are high enough that they easily qualify to be broad

resonances. This implies that the narrow resonance formula used to see their effect is not

precise, but still gives a rough idea of whether the resonance is ineffective or not.

Our expression for thermonuclear rates in the narrow resonance approximation is ex-

plained in detail in Appendix A.1, and is given by

〈σv〉 = ω Γeff

(

µT

)3/2

e−|Eres|/T f (2Eres/Γtot) (2.13)

This rate is controlled by two parameters specific to the compound nuclear state: Eres and

Γeff . Here Eres is given in eq. (2.8), and measures the offset from the entrance channel and

27

the compound state. The resonance strength is quantified via

Γeff =ΓinitΓfin

Γtot(2.14)

with Γinit and Γfin being the entrance and exit widths of a particular reaction, and Γtot the

sum of the widths of all possible channels. Of these widths, the smaller of Γinit and Γfin along

with Γtot are listed in the table above. The resonance strength, Γeff ≈ Γinit, if Γfin dominates

the total width, and vice versa. If Γinit and Γfin are the dominant partial widths and they

are comparable to each other, then the strength is even higher.

As discussed in Appendix A.1, our narrow resonance rate in eq. (2.13) improves on

the form of the usual expression for narrow resonance in two ways: (a) it extends to the

subthresold domain; and (b) it introduces the factor f which accounts for a finite Eres/Γtot

ratio.

It is important to make a systematic and comprehensive search for all possible experimen-

tally identified resonances capable of removing this discrepancy. In addition, it is possible

that resonances and indeed energy levels themselves were missed, especially at the higher

energies, where uncertainties are greater. Therefore it is useful to map the parameter space

where the lithium discrepancy is removed to apriori lay down our expectations of such missed

resonances. This can be done by looking at interesting initial states involving 7Li and 7Be,

and abundant projectiles p, n, d, t, 3He, α, and parametrizing the effect of inelastic channels

on the mass-7 abundance. This is described in § 2.4.

2.4 Narrow resonance solution space

In order to study the effect of resonances in different compound nuclei on the abundance of

mass 7, our strategy is to specify the reaction rate for possible resonances, and then run the

BBN code to find the mass-7 abundance in the presence of these resonances. In particular,

for reactions involving light projectile X, we are interested in considering the general effect

28

of states 7A+X → C∗, including those associated with known energy levels in the compound

nucleus, as well as possible overlooked states.

We assume that the narrow resonance approximation holds true at least as a rough guide.

If the reaction pathway is specified, i.e., all of the nuclei 7A+X → C∗ → Y +Z are identified,

then the reduced mass µ, reverse ratio and the Q-value are uniquely determined. In this

case, the thermally averaged cross-section is given by eq. (2.13), with two free parameters:

the product ωΓeff and the resonance energy, Eres. Because the state C∗ is unspecified, so is

its spin J∗. On the other hand, we do know the spins of the initial state particles, and thus

ω is specified up to a factor 2J∗ +1 (eq. 2.10). For this reason, the ωΓeff dependence reduces

to (2J∗ + 1)Γeff , which we explicitly indicate in all of our plots.

In a few cases we will be interested in one specific final quantum state, e.g., 7Be(t, 3He)7Li;

when the final state is specified, the reaction can be completely determined, including the

effect of the reverse rates. However, in most situations we are interested in the possibility

of an overlooked excited state in the compound nucleus, and thus in unknown final states.

In this scenario we thus have only a “generic” inelastic exit channel. Consequently, for such

plots we cannot evaluate the reverse reaction rate (which is in all interesting cases small)

and so we set the reverse ratio to zero.

The resonant rates are included in the BBN code, individually for compound nuclei with

an interesting initial state. The plots below show contours of constant, reduced mass-7

abundances. A general feature of all the plots, is the near linear relation between log Γeff

and Eres in the region of larger, positive values of Γeff and Eres. This can seen quantitatively

as follows. The thermal rate is integrated over time or equivalently temperature to give the

final abundance of mass 7 or 7Li as it exists. Now assuming that the thermal rate operates at

an effective temperature, TLi, at which 7Li production peaks, a given value for this effective

〈σv〉 will give a fixed abundance. This implies,

δY7/Y7 ∼ 〈σv〉peak ∼ Γeff e−Eres/TLi ∼ constant (2.15)

29

This gives a feel for the linear relation in the plot.

2.4.1 A = 8 Compound Nucleus

Compound Nucleus, Initial Linit Lfin Eres Γtot Exit Exit ChannelJπ, Eex State Channels Width

8Li, 3+, 2.255 MeV 7Li + n 1 1 222.71 keV 33 ± 6 keV γ(ground state) 7.0 ± 3.0 × 10−2 eV(Included) 1 n (elastic) ≈100% 33 ± 6 keV

8Be, 2+, 16.922 MeV 7Li + p 1 2 -333.1 keV 74.0 ± 0.4 keV γ(ground state) 8.4 ± 1.4 × 10−2 eV1 γ(3.04 MeV) < 2.80 ± 0.18 eV2 α ≈ 100% ≈ 74.0 keV1 p (elastic) unknown

8Be, 1+, 17.640 MeV 7Li + p 1 1 384.9 keV 10.7 keV γ(ground state) 16.7 eV1 γ(3.04 MeV) 6.7 ± 1.3 eV2 γ(3.04 MeV) 0.12 ± 0.05 eV1 γ(16.63 MeV) (3.2 ± 0.3) × 10−2 eV1 γ(16.92 MeV) (1.3 ± 0.3) × 10−3 eV1 p (elastic) 98.8% 10.57 keV

8Be, 2−, 18.91 MeV 7Be + n 0 1 10.3 keV 122 keV* γ(16.922 MeV) 9.9 ± 4.3 × 10−2 eV(Included) 1 γ (16.626 MeV) 0.17 ± 0.07 eV

0 p < 105.1 keV*2 p+ 7Li

∗(0.4776 MeV) < 105.1 keV*

0 n (elastic) 16.65 keV*8Be, 3+, 19.07 MeV 7Be + n 1 1 170.3 keV 270 ± 20 keV p ≈ 100% < 270 keV

(Included) 3 p+ 7Li∗

(0.4776 MeV) < 270 keV1 γ(3.03 MeV) 10.5 eV1 n (elastic) unknown

8Be, 3+, 19.235 MeV 7Be + n 1 1 335.3 keV 227 ± 16 keV p ≈ 50% ≈ 113.5 keV(Included) 1 γ(3.03 MeV) 10.5 eV

1 n (elastic) ≈ 50% ≈ 113.5 keV8Be, 1−, 19.40 MeV 7Be + n 0 0 500.3 keV 645 keV p unknown

0 p+ 7Li∗

(0.4776 MeV) unknown0 n (elastic) unknown1 α unknown

8Bg.s.

, 2+, 0 MeV 7Be + p 1 1 -0.1375 MeV unknown p (elastic) unknown0 EC→ 8Be 8.5 × 10−19 eV

8B, 1+, 0.7695 MeV 7Be + p 1 1 630 ±3 keV 35.7 ± 0.6 keV γ (ground state) 25.2 ± 1.1 meV(Included) 1 p (elastic) 100% 35.7 ± 0.6 keV

Table 2.1: This table lists the potential resonances in 8Li, 8Be and 8B which may achieverequired destruction of mass 7. These are all allowed by selection rules and includes someresonances already accounted for in determining the current theoretical 7Li abundance in-dicated as (Included). The entrance and exit channels along with their partial and totalwidths (Γtot), minimum angular momenta (Linit, Lfin) as well as resonance energies are listedwherever experimental data are available. The starred widths are a result of fits from R-matrix analysis. The list includes final products in ground and excited states with the lattermarked with a star in the superscript.

As seen in Table 2.1, the only resonance energy level of interest in the 8Li compound

nucleus at 2.255 MeV is already accounted for in the 7Li+n reaction. In the 8Be compound

nucleus, there are six levels of relevance for destroying either 7Li or 7Be at 16.922, 17.64,

18.91, 19.07, 19.24 and 19.40 MeV within our limit on Eres. The 16.922 MeV level is more

30

than 300 keV below threshold and has a maximum total width of only 74 keV. Therefore,

it is expected to have a weak effect. The 17.64 MeV level has typically low photon widths

(≈ 20 eV ) and a total width of 10.7 keV. But this state’s decay is dominated by the elastic

channel which makes this channel uninteresting.

The energy level diagram for 8Be 7(Tilley et al., 2004) shows the initial state, 7Be + n at

an entrance energy of E = 18.8997 MeV bringing the 18.91, 19.07, 19.235 and 19.40 MeV

levels into play. From among these the effect of the 18.91, 19.07 and 19.235 MeV resonances

are already accounted for in the well known 7Be(n, p)7Li reaction Cyburt (2004a). The

(18.91 MeV, 2−), resonance with Linit = 0 is the dominant contributor Coc et al. (2004b);

Adahchour & Descouvemont (2003). Being a broad resonance with a total width of ≈ 122

keV, the Breit-Wigner form is not used and instead an R-matrix fit to the data Cyburt

(2004a), is used to evaluate the contribution of the resonant rate. The remaining level

at 19.40 should also contribute to this reaction through ground and excited states. Only

the 19.40 MeV channel can have an α exit channel due to parity considerations. And this

resonance, despite a high resonance energy of ≈ 500 keV, can in principle be important due

to its large total width of 645 keV, if the proton branching ratio is high.

Figure 2.1 shows the 7Li abundance in the (Γeff , Eres) plane for the 7Be(n, p)7Li reaction.

Contours for 7Li/H ×1010 = 1.23, 2.0, 3.0, 4.0, and 5.0 (as labelled) are plotted as functions

of the effective width and resonant energy. Below ≈ (2J∗ + 1)40 = 120 keV, we expect

our results based on the narrow resonance approximation to be quite accurate. As one can

see from this figure, to bring the 7Li abundance down close to observed values, one would

require a very low resonance energy (of order ±30 keV) with a relatively large effective width.

Unfortunately, the 19.40 MeV level of 8Be corresponds to Eres = 500 keV as shown by the

vertical dashed line and does not make any real impact on the 7Li abundance.

Figure 2.2 shows the effect of a 8B resonance with 7Be and p in the initial state, plotted

in the (Γeff , Eres) plane with contours of constant mass 7 abundances. According to Fig. 2.2

7http : //www.tunl.duke.edu/nucldata/figures/08figs/08 04 2004.gif.

31

Figure 2.1: The effect of resonances in the 8Be compound nucleus involving initial states7Be+n. It shows the range of values for the product of the resonant state spin degeneracyand resonance strength, (2J + 1)Γeff , versus the resonance energy. Contours indicate wherethe lithium abundance is reduced to 7Li/H = 1.23×10−10 (red), 2.0×10−10 (blue), 3.0×10−10

(green) 4.0 × 10−10 (black) and 5.0 × 10−10 (magenta). Normal resonances have Eres > 0,while subthreshold resonances lie in the Eres < 0. The horizontal dot-dashed black line is theexperimental value of the strength of the resonance corresponding to the 19.40 MeV energylevel. The vertical dashed black line shows the position of Eres for the same state.

32

for resonance energies of a few 10’s of keV’s, resonance strength of a few meV is sufficient

to attain the observational value of mass 7. However, from the energy level diagram for 8B,

8(Tilley et al., 2004), the closest resonant energy level, Eex is at 0.7695 MeV 9(Tilley et al.,

2004), whose effect is already included via the 7Be(p, γ)8B reaction. The experimental value

of resonance energy is 632 keV which is off the scale in this figure. The only other close

energy level to the 7Be+ p entrance channel is at -0.1375 MeV which means that the ground

state is a sub-threshold state. This is not the usual resonant reaction, since the ground state

doesn’t have a width in the sense we refer to a width for the other reactions. But at these

energies, the astrophysical S-factor is ≈ 10 eV-barn which is very small and will yield a low

cross-section. This too is off scale in the figure and verifies that the 7Be(p, γ)8B reaction

does not yield an important destruction channel.

2.4.2 A = 9 Compound Nucleus

The energy level diagram for 9Be, 10(Tilley et al., 2004) shows energy levels of interest at

16.671, 16.9752 and 17.298 MeV; these appear in Table 2.2. The 7Li + d entrance channel

sits at 16.6959 MeV. The lowest lying resonant state is at 16.671 MeV and is a sub-threshold

state with Eres = −24.9 keV which lies within the total width of 41 keV. This resonance

is thus obviously tantalizing–it is well-tuned energetically and involves an abundant, stable

projectile. The 7Li abundance contours for the 9Be resonance states are shown in Fig. 2.3.

Perhaps disappointingly, the figure shows that the effect on primordial mass 7 is minor. This

illustrates the inability of direct 7Li destruction channels to reduce the mass-7 abundance,

as explained in §2.2. Given the overall difficulty of this channel, it is clear that the other

possible resonant energy levels (16.9752 MeV and 17.298 MeV) also fail to substantially

reduce the mass-7 abundance.

The 9B compound nucleus is relevant for studying the effect of the 7Be(d, p)2α and its

8http : //www.tunl.duke.edu/nucldata/figures/08figs/08 05 2004.gif.9http : //www.tunl.duke.edu/nucldata/figures/08figs/08 05 2004.gif.

10http : //www.tunl.duke.edu/nucldata/figures/09figs/09 04 2004.gif.

33

Figure 2.2: As in Fig. 2.1 for the resonances in the 8B compound nucleus involving initialstates 7Be+p.

34

Figure 2.3: As in Fig. 2.1 for the resonances in the 9Be compound nucleus.

35

Compound Nucleus, Initial Linit Lfin Eres Γtot Exit Exit ChannelJπ, Eex State Channels Width

9Be, 7Li + d 1 unknown -24.9 keV 41 ± 4 keV γ unknown(5/2+), 16.671 MeV 2 n+ 8Be unknown

0 n+ 8Be∗

(3.03 MeV) unknown2 n+ 8Be

∗(11.35 MeV) unknown

0 p unknown1 α unknown1 d (elastic) unknown

9Be, 7Li + d 0 1 279.3 keV 389 ± 10 eV γ (ground state) 16.9 ± 1.0 eV1/2−, 16.9752 MeV 1 γ (1.68 MeV) 1.99 ± 0.15 eV

2 γ (2.43 MeV) 0.56 ± 0.12 eV1 γ (2.78 MeV) 2.2 ± 0.7 eV

unknown γ (Unknown level TUNL) < 0.8 eV1 γ (4.70 MeV) 2.2 ± 0.3 eV1 p 12+12

−6 eV1 n < 288 eV1 n+ 8Be

∗(3.03 MeV) < 288 eV

3 n+ 8Be∗

(11.35 MeV) < 288 eV2 α < 241 eV0 d (elastic) 62 ± 10 eV

9Be, 7Li + d 0 unknown 602.1 keV 200 keV γ (ground state) unknown(5/2)−, 17.298 MeV 1 p 194.4 keV

(Included) 3 n+ 8Be unknown1 n+ 8Be

∗(3.03 MeV) unknown

1 n+ 8Be∗

(11.35 MeV) unknown2 α unknown0 d (elastic) unknown

Table 2.2: As in Table 2.1 listing the potential resonances in 9Be.

Compound Nucleus, Initial Linit Lfin Eres Γtot Exit Exit ChannelJπ, Eex State Channels Width

9B, (5/2+), 16.71 MeV 7Be + d 1 2 219.9 keV unknown p+ 8Be unknown0 p+ 8Be

∗(3.03 MeV) unknown

2 p+ 8Be∗

(11.35 MeV) unknown0 p+ 8Be

∗(16.626 MeV) unknown

0 p+ 8Be∗

(16.922 MeV) unknown2 3He unknown1 α + 5Li unknown3 α + 5Li

∗(1.49 MeV) unknown

1 d (elastic) unknown9B, (1/2−), 17.076 MeV 7Be + d 0 1 585.9 keV 22 keV p+ 8Be unknown

1 p+ 8Be∗

(3.03 MeV) unknown3 p+ 8Be

∗(11.35 MeV) unknown

1 p+ 8Be∗

(16.626 MeV) unknown1 3He unknown2 α + 5Li unknown0 α + 5Li

∗(1.49 MeV) unknown

0 d (elastic) unknown

Table 2.3: As in Table 2.1 listing the potential resonances in 9B.

competitors such as 7Be(d, 3He)6Li and 7Be(d, α)5Li. As seen in Table 2.3, the only two

levels of interest here are the 16.71 and 17.076 MeV levels. The 16.71 MeV level corresponds

to a resonance energy of 220 keV as shown by the vertical dashed line in Fig. 2.4 11(Tilley

et al., 2004). The widths are unknown experimentally. The approximate narrow resonance

11http : //www.tunl.duke.edu/nucldata/figures/09figs/09 05 2004.gif.

36

limit on the resonance width which is shown by the horizontal solid line is around 40 keV.

The p exit channel leads to the 7Be(d, p)8Be∗

reaction through the excited state at 16.63

MeV in 8Be. This should eventually lead to formation of alpha particles. Fig. 2.4 shows the

effect of the 16.71 MeV resonance on the mass-7 abundance as a function of the resonance

strength and energy under the narrow resonance approximation. From the plot, we see that

the 7Li abundance is reduced by 50% for (2J + 1)Γeff = 240 keV. This state has J = 5/2

and therefore, a value Γeff = 40 keV or more will have substantial impact on the problem.

Furthermore, as ΓL ≥ Γeff , we require ΓL >∼ 40 keV. This result confirms the conclusion of

Cyburt & Pospelov (2012). Later in § 2.5 we will see how this compares with theoretical

limits. As the decay widths are largely unknown, experimental data on the width are needed.

The state at 17.076 MeV corresponds to a resonant energy of Eres = 586 keV and is

beyond the scale shown in Fig. 2.4. A solution using this state is very unlikely.

2.4.3 A = 10 Compound Nucleus

Table 2.4 shows that the 10Be compound nucleus has energy levels at 17.12 and 17.79 MeV

12(Tilley et al., 2004) which are close to the initial state 7Li + t at 17.2509 MeV. The former

is far below threshold and does not contribute to 7Li destruction. The 17.79 MeV level is

around 540 keV above the entrance energy and its spin and parity are unknown. The total

width 13(Tilley et al., 2004) is Γtot = 112 keV which implies a small overlap with the entry

channel which renders this resonance insignificant despite having a number of n exit channels

with both ground state and excited states of 9Be. As seen in Fig. 2.5, the effect of 7Li + t is

small for the interesting region of parameter space.

The 10B compound nucleus has energy levels at 18.2, 18.43, 18.80 and 19.29 MeV, which

we investigate. The 18.2 MeV level is uncertain experimentally 14(Tilley et al., 2004) as

indicated in Table 2.5, and hence ideal for parametrizing. There is a 3He entrance channel a

12http : //www.tunl.duke.edu/nucldata/figures/10figs/10 04 2004.gif.13http : //www.tunl.duke.edu/nucldata/figures/10figs/10 04 2004.gif.14http : //www.tunl.duke.edu/nucldata/figures/10figs/10 05 2004.gif.

37

Figure 2.4: As in Fig. 2.1 for the resonances in the 9B compound nucleus. The verticaldashed line at 220 keV indicates the experimental central value of the resonance energy ofthe 16.71 MeV level.

38

Compound Nucleus, Initial State Linit Lfin Eres Γtot Exit Exit ChannelJπ, Eex Channels Width10Be, 7Li + t 0 0 -130.9 keV ≈ 150 keV n+ 9Be unknown

(2−), 17.12 MeV 1 n+ 9Be∗

(1.684 MeV) unknown0 n + 9Be

∗(2.4294 MeV) unknown

2 n+ 9Be∗

(2.78 MeV) unknown1 n+ 9Be

∗(3.049 MeV) unknown

1 n+ 9Be∗

(4.704 MeV) unknown0 n+ 9Be

∗(5.59 MeV) unknown

2 n+ 9Be∗

(6.38 MeV) unknown3 n+ 9Be

∗(6.76 MeV) unknown

0 n+ 9Be∗

(7.94 MeV) unknown0 t (elastic) unknown

10Be, 7Li + t unknown unknown 539.1 keV 112 ± 35 keV γ 3 + 2 eVunknown, 17.79 MeV unknown n+ 9Be < 77 keV

unknown n+ 9Be∗

(1.684 MeV) < 77 keVunknown n + 9Be

∗(2.4294 MeV) < 77 keV

unknown n+ 9Be∗

(2.78 MeV) < 77 keVunknown n+ 9Be

∗(3.049 MeV) < 77 keV

unknown n+ 9Be∗

(4.704 MeV) < 77 keVunknown n+ 9Be

∗(5.59 MeV) < 77 keV

unknown n+ 9Be∗

(6.38 MeV) < 77 keVunknown n+ 9Be

∗(6.76 MeV) < 77 keV

unknown n+ 9Be∗

(7.94 MeV) < 77 keVunknown t (elastic) 78 keV

Table 2.4: As in Table 2.1 listing the potential resonant reactions in 10Be.

Compound Nucleus, Initial Linit Lfin Eres Γtot Exit Exit ChannelJπ, Eex State Channels Width

10B, 7Li + 3He unknown unknown 411.7 keV 1500 ± 300 keV p+ 9Be unknownunknown, (18.2 MeV) unknown p+ 9Be

∗(1.684 MeV) unknown

unknown p+ 9Be∗

(2.4294 MeV) unknownunknown p+ 9Be

∗(2.78 MeV) unknown

unknown p+ 9Be∗

(3.049 MeV) unknownunknown p+ 9Be

∗(4.704 MeV) unknown

unknown p+ 9Be∗

(5.59 MeV) unknownunknown p+ 9Be

∗(6.38 MeV) unknown

unknown p+ 9Be∗

(6.76 MeV) unknownunknown p+ 9Be

∗(7.94 MeV) unknown

unknown p+ 9Be∗

(11.283 MeV) unknownunknown n+ 9B unknownunknown n + 9B

∗(1.6 MeV) unknown

unknown n + 9B∗

(2.361 MeV) unknownunknown n+ 9B

∗(2.75 MeV) unknown

unknown n + 9B∗

(2.788 MeV) unknownunknown n + 9B

∗(4.3 MeV) unknown

unknown n+ 9B∗

(6.97 MeV) unknownunknown d+ 8Be unknownunknown d+ 8Be

∗(3.03 MeV) unknown

unknown d+ 8Be∗

(11.35 MeV) unknownunknown t unknownunknown α + 6Li unknownunknown α+ 6Li

∗(2.186 MeV) unknown

unknown α+ 6Li∗

(3.563 MeV) unknownunknown α + 6Li

∗(4.31 MeV) unknown

unknown α + 6Li∗

(5.37 MeV) unknownunknown 3He (elastic) unknown

Table 2.5: As in Table 2.1 listing the ground and excited final states for the 18.2 MeV energylevel in 10B.

39

Figure 2.5: As in Fig. 2.1 for the resonances in the 10Be compound nucleus involving theinitial state 7Li+t (left), and in the 10B compound nucleus involving the initial state 7Li+3He.(right).

Compound Nucleus, Initial Linit Lfin Eres Γtot Exit Exit ChannelJπ, Eex State Channels Width

10B, 7Li + 3He 0 unknown 641.7 keV 340 keV γ(ground state) ≥ 3 eV2−, 18.43 MeV unknown γ (4.77 MeV) ≥ 17eV

0 n+ 9B unknownunknown n + 9B

∗(1.6 MeV) unknown

0 n+ 9B∗

(2.361 MeV) unknown2 n+ 9B

∗(2.75 MeV) unknown

1 n+ 9B∗

(2.788 MeV) unknownunknown n + 9B

∗(4.3 MeV) unknown

2 n+ 9B∗

(6.97 MeV) unknown0 p + 9Be unknown1 p+ 9Be

∗(1.684 MeV) unknown

0 p+ 9Be∗

(2.4294 MeV) unknown2 p+ 9Be

∗(2.78 MeV) unknown

1 p+ 9Be∗

(3.049 MeV) unknown1 p+ 9Be

∗(4.704 MeV) unknown

0 p+ 9Be∗

(5.59 MeV) unknown2 p+ 9Be

∗(6.38 MeV) unknown

3 p+ 9Be∗

(6.76 MeV) unknown0 p+ 9Be

∗(7.94 MeV) unknown

2 p+ 9Be∗

(11.283 MeV) unknown0 p+ 9Be

∗(11.81 MeV) unknown

1 d+ 8Be unknown1 d+ 8Be

∗(3.03 MeV) unknown

1 d+ 8Be∗

(11.35 MeV) unknown1 α + 6Li unknown1 α + 6Li

∗(2.186 MeV) unknown

1 α + 6Li∗

(4.31 MeV) unknown1 α + 6Li

∗(5.37 MeV) unknown

1 α + 6Li∗

(5.65 MeV) unknown0 3He (elastic) unknown

Table 2.6: As in Table 2.1 listing the ground and excited final state channels for the 18.43MeV energy level in 10B for the 7Li + 3He initial state.

little over 400 keV below this level. The current total experimental width is 1.5 MeV which is

very large and the branching ratios are unknown. The current uncertainty in the Eres is 200

40

Compound Nucleus, Initial Linit Lfin Eres Γtot Exit Exit ChannelJπ, Eex State Channels Width

10B, 7Be + t 0 unknown -239.1 keV 340 keV γ(ground state) ≥ 3 eV2−, 18.43 MeV unknown γ (4.77 MeV) ≥ 17 eV

0 n + 9B unknownunknown n+ 9B

∗(1.6 MeV) unknown

0 n+ 9B∗

(2.361 MeV) unknown2 n + 9B

∗(2.75 MeV) unknown

1 n+ 9B∗

(2.788 MeV) unknownunknown n+ 9B

∗(4.3 MeV) unknown

2 n + 9B∗

(6.97 MeV) unknown0 p + 9Be unknown1 p+ 9Be

∗(1.684 MeV) unknown

0 p+ 9Be∗

(2.4294 MeV) unknown2 p+ 9Be

∗(2.78 MeV) unknown

1 p+ 9Be∗

(3.049 MeV) unknown1 p+ 9Be

∗(4.704 MeV) unknown

0 p+ 9Be∗

(5.59 MeV) unknown2 p+ 9Be

∗(6.38 MeV) unknown

3 p+ 9Be∗

(6.76 MeV) unknown0 p+ 9Be

∗(7.94 MeV) unknown

2 p+ 9Be∗

(11.283 MeV) unknown0 p+ 9Be

∗(11.81 MeV) unknown

1 d+ 8Be unknown1 d+ 8Be

∗(3.03 MeV) unknown

1 d+ 8Be∗

(11.35 MeV) unknown1 α+ 6Li unknown1 α + 6Li

∗( 2.186 MeV) unknown

1 α + 6Li∗

( 4.31 MeV) unknown1 α + 6Li

∗( 5.37 MeV) unknown

1 α + 6Li∗

( 5.65 MeV) unknown0 t (elastic) unknown

Table 2.7: As in Table 2.1 listing the ground and excited final state channels for the 18.43MeV energy level in 10B for the 7Be + t initial state.

keV. However, according to the plot in Fig. 2.5, even a 200 keV reduction in Eres would not

be sufficient to cause any appreciable destruction of 7Li as this reaction has negligible effect

on the mass-7 abundance. This is another illustration of the fact that reactions involving

direct destruction of 7Li are unimportant.

The 18.43 MeV level is better understood Yan et al. (2002) and with a resonance energy

of ∼ 640 keV for the 7Li + 3He initial state (Table 2.6) and a total width of 340 keV has

a lower entrance probability and therefore is likely to be ineffective. This is evident from

Fig. 2.5. This level is also a sub-threshold resonance for the 7Be + t state (Table 2.7), with

resonance energy, Eres = −239.1 keV. This is far below threshold rendering it ineffective.

Staying with 7Be+ t, the closest energy level above the entrance energy of 18.669 MeV is

the 18.80 MeV (2+) level (Table 2.8), which corresponds to a resonance energy of ≈ 130 keV

41

Compound Nucleus, Initial State Linit Lfin Eres Γtot Exit Exit ChannelJπ, Eex Channels Width

10B, 7Be + t 1 unknown 130.9 keV < 600 keV γ (0.72 MeV) ≥ 20 eV2+, 18.80 MeV unknown γ (3.59 MeV) ≥ 20 eV

1 n + 9B unknownunknown n+ 9B

∗(1.6 MeV) unknown

1 n + 9B∗

(2.361 MeV) unknown1 n+ 9B

∗(2.75 MeV) unknown

0 n + 9B∗

(2.788 MeV) unknownunknown n+ 9B

∗(4.3 MeV) unknown

1 n+ 9B∗

(6.97 MeV) unknown1 p + 9Be unknown2 p+ 9Be

∗(1.684 MeV) unknown

1 p+ 9Be∗

(2.4294 MeV) unknown1 p+ 9Be

∗(2.78 MeV) unknown

0 p+ 9Be∗

(3.049 MeV) unknown0 p+ 9Be

∗(4.704 MeV) unknown

1 p+ 9Be∗

(5.59 MeV) unknown1 p+ 9Be

∗(6.38 MeV) unknown

2 p+ 9Be∗

(6.76 MeV) unknown1 p+ 9Be

∗(7.94 MeV) unknown

1 p+ 9Be∗

(11.283 MeV) unknown1 p+ 9Be

∗(11.81 MeV) unknown

2 d+ 8Be unknown0 d+ 8Be

∗(3.03 MeV) unknown

2 d+ 8Be∗

(11.35 MeV) unknown1 3He + 7Li unknown1 3He + 7Li

∗(0.47761 MeV) unknown

2 α + 6Li unknown2 α+ 6Li

∗(2.186 MeV) unknown

2 α + 6Li∗

(3.56 MeV) unknown0 α + 6Li

∗(4.31 MeV) unknown

0 α + 6Li∗

(5.37 MeV) unknown2 α + 6Li

∗(5.65 MeV) unknown

1 t (elastic) unknown

Table 2.8: As in Table 2.1 listing the ground and excited final state channels for the 18.80MeV energy level in 10B for the 7Be + t initial state.

15(Tilley et al., 2004). The exit channel widths for p and 3He are unknown experimentally and

thus, this is a candidate for parametrization. There is a weak upper limit on Γtot < 600 keV

16(Tilley et al., 2004), which for J∗ = 2 is off scale in Figs. 2.6 and 2.7. The contour plot

in Fig. 2.6 show that for a central value of resonance energy of ≈ 130 keV shown by the

vertical dashed line, resonance strength of just under 1 MeV is required which is very high.

Also, parity requirements force L = 1, which will cause suppression of this channel. We note

that there is no quoted uncertainty for this energy level and neighboring levels have typical

uncertainties of 100-200 keV. Therefore it may be possible (within 1-2σ) that the state lies

at an energy of 100 keV lower and would energetically, have a chance at solving the 7Li

15http : //www.tunl.duke.edu/nucldata/figures/10figs/10 05 2004.gif.16http : //www.tunl.duke.edu/nucldata/figures/10figs/10 05 2004.gif.

42

Compound Nucleus, Initial State Linit Lfin Eres Γtot Exit Exit ChannelJπ, Eex Channels Width

10B, 2−, 19.29 MeV 7Be + t 0 unknown 620.9 keV 190 ± 20 keV γ unknown0 n + 9B unknown

unknown n+ 9B∗

(1.6 MeV) unknown0 n + 9B

∗(2.361 MeV) unknown

2 n+ 9B∗

(2.75 MeV) unknown1 n + 9B

∗(2.788 MeV) unknown

unknown n+ 9B∗

(4.3 MeV) unknown0 p + 9Be unknown1 p+ 9Be

∗(1.684 MeV) unknown

0 p+ 9Be∗

(2.4294 MeV) unknown2 p+ 9Be

∗(2.78 MeV) unknown

1 p+ 9Be∗

(3.049 MeV) unknown1 p+ 9Be

∗(4.704 MeV) unknown

0 p+ 9Be∗

(5.59 MeV) unknown2 p+ 9Be

∗(6.38 MeV) unknown

3 p+ 9Be∗

(6.76 MeV) unknown0 p+ 9Be

∗(7.94 MeV) unknown

2 p+ 9Be∗

(11.283 MeV) unknown0 p+ 9Be

∗(11.81 MeV) unknown

1 d+ 8Be unknown1 d+ 8Be

∗(3.03 MeV) unknown

3 d+ 8Be∗

(11.35 MeV) unknown0 3He unknown1 α + 6Li unknown1 α+ 6Li

∗(2.186 MeV) unknown

1 α + 6Li∗

(4.31 MeV) unknown1 α + 6Li

∗(5.37 MeV) unknown

1 α + 6Li∗

(5.65 MeV) unknown0 t (elastic) unknown

Table 2.9: As in Table 2.1 listing the ground and excited final state channels for the 19.29MeV energy level in 10B for the 7Be + t initial state.

problem. This is true for the p exit channel.

The 3He exit channel may also reduce mass 7, through the formation of the 7Li which is

much easier to destroy. This is reflected in Fig. 2.7, which shows that at resonance energies

of ≤ 100 keV, a strength of a few 100 keV, but less than 600 keV may be sufficient to achieve

comparable destruction of 7Be as the 16.71 MeV resonance. The caveat is that for such values

of strengths, the narrow resonance approximation does not hold true and this may lead to a

reduced effect. Nevertheless, this is yet another case deserving a detailed comparison with

the theoretical limits which will follow in § 2.5. Once again, definitive conclusions can be

drawn only based on experimental data.

The 19.29 MeV level (Table 2.9) is energetically harder to access and with a total width

of only 190 keV, it is unlikely to be of significance, despite being less studied.

43

Figure 2.6: As in Fig. 2.1 for the resonances in the 10B compound nucleus involving initialstates 7Be+t.

44

Figure 2.7: As in Fig. 2.1 for the resonances in the reaction 7Be(t,3He)7Li.

45

Compound Nucleus, Initial Linit Lfin Eres Γtot Exit Exit ChannelJπ, Eex State Channels Width

10C, 7Be + 3He unknown unknown unknown unknown p unknownunknown unknown (Q = 15.003 MeV) α unknown

3He (elastic) unknown11B, 7Li + α 0 1 -103.7 keV 1.346 eV γ (ground state) 0.53 ± 0.05 eV

(3/2−), 8.56 MeV 1 γ (2.125 MeV) 0.28 ± 0.03 eV1 γ (4.445 MeV) (4.7 ± 1.1) × 10−2 eV1 γ (5.020 MeV) (8.5 ± 1.2) × 10−2 eV1 α (elastic) Unknown

11B, 7Li + α 2 1 256.3 keV 4.37 ± 0.02 eV γ (ground state) 4.10 ± 0.20 eV(5/2−), 8.92 MeV 2 γ (ground state) (5.0 ± 3.6) × 10−2 eV

(Included) 1 γ (4.445 MeV) 0.22 ± 0.02 eV1 α (elastic) Unknown

11B, 7Li + α 3 1 526.3 keV 1.9+1.5−1.1 eV γ (ground state) (2.7 ± 1.2) × 10−3eV

7/2+, 9.19 MeV 2 γ (4.445 MeV) 0.25 ± 0.09 eV0 γ (6.743 MeV) (3.8 ± 1.3) × 10−2 eV1 α (elastic) Unknown

11B, 7Li + α 1 1 606.3 keV 4 keV γ (ground state) 0.212 eV5/2+, 9.271 MeV 0 γ (4.445 MeV) 0.802 eV

0 γ (6.743 MeV) 0.137 eV1 γ (6.792 MeV) < 0.007 eV1 α (elastic) ≈ 4 keV

11C, 7Be + α 1 1 -43.3 keV 0.0105 eV γ (ground state) Unknown3/2+, 7.4997 MeV 0 γ (2.0 MeV) Unknown

1 α(elastic) Unknown11C, 7Be + α 0 1 557 keV 11 ± 7 eV γ (ground state) 0.26 ± 0.06 eV

(3/2−), 8.10 MeV 1 γ (2.0 MeV) (9.1 ± 2.3) × 10−2 eV0 α(elastic) Unknown

Table 2.10: As in Table 2.1 listing resonances in 10C, 11B and 11C.

The 10C nucleus 17(Ajzenberg-Selove, 1990) appearing in Table 2.10 shows large uncer-

tainties and experimental gaps at higher energy levels which may be relevant to entrance

channels involving 7Be. Reactions involving the 7Be + 3He initial state could contribute in

destroying 7Be if there exists a resonance in the parameter space shown in the Fig. 2.8.

These reactions win over those involving the 7Be+ t state, because 3He is substantially more

abundant than t, but are worse off due to a higher Coulomb barrier. The entrance energy

for 7Be + 3He is 15.0 MeV. As one can see from the figure, a 1− or 2− state with a resonance

energy of either -10 keV or 40 keV corresponding to energy levels of 14.99 and 15.04 MeV

respectively with a strength as high as a few 10’s of keVs is what it will take to solve the

lithium problem with this initial state. Thus, any 10C resonance near these energies which

may have been missed by experiment may be interesting as a solution to the lithium problem;

we return to this issue in more detail in § 2.5.

17http : //www.tunl.duke.edu/nucldata/figures/11figs/11 06 1990.gif.

46

Figure 2.8: As in Fig. 2.1 for the resonances in 10C involving initial state 7Be + 3He

2.4.4 A = 11 Compound Nucleus

For 11B 18(Ajzenberg-Selove, 1990), Table 2.10 shows that the entrance channel, 7Li+α is at

8.6637 MeV which is 103.7 keV above the resonant energy level at 8.560 MeV and ≈ 260 keV

below the resonant energy level at 8.92 MeV. Parity demands angular momentum to be 0.

Both states are at relatively large |Eres| and are not capable of making a sizable impact on

the 7Li abundance. Table 2.10 further lists states 9.19 MeV (which requires L = 3 and has

a total width of < 2 eV) and 9.271 MeV (whose decay is dominated by the elastic channel)

18http : //www.tunl.duke.edu/nucldata/figures/11figs/11 05 1990.gif.

47

which have progressively larger resonant energies and are unlikely to provide a solution.

For 11C 19(Ajzenberg-Selove, 1990), the entrance channel, 7Be+α is at 7.543 MeV which

is 43 keV above the resonant energy level at 8.560 MeV and 557 keV below the resonant

energy level at 8.10 MeV.

Figure 2.9: As in Fig. 2.1 for the resonances in 11C involving initial states 7Be+α.

As seen in Fig. 2.9, we find that the sub-threshold resonance in the 11C nucleus, produces

a very insignificant effect on 7Be in agreement with the claim in Cyburt & Pospelov (2012).

The super-threshold resonance states are also too far away at resonance energies, 557 keV

19http : //www.tunl.duke.edu/nucldata/figures/11figs/11 06 1990.gif.

48

and 260 keV for 7Be(α, γ)11C and 7Li(α, γ)11B respectively.

However, Fig. 2.9 shows that the presence of a (missed) resonance at resonance energies

of few tens of keV ’s, requires a very meagre strength of the order of tens of meV ’s to destroy

mass 7 substantially. Strengths of this order are typical of electromagnetic channels. It is

difficult to assess the probability that a 11C state at 7.55 MeV has been overlooked.

2.5 Reduced List of Candidate Resonances

Having systematically identified all possible known resonant energy levels which could affect

BBN, we find most of these levels are ruled out immediately as promising solutions, based on

their measured locations, strengths, and widths. As expected, the existing electromagnetic

channels are too weak to cause significant depletion of lithium owing to their small widths.

From amongst the various hadronic channels listed in the tables above, we have seen

that all channels are unimportant except three which are summarized in table 7.1. The

7Be+d channels involving the 16.71 MeV resonance in 9B, the 7Be+ t channels involving the

18.80 MeV resonance in 10B and 7Be + 3He channels. These are ones where a more detailed

theoretical calculation of widths is required to decide whether they may be important or not.

For each reaction, the Wigner limit, eq. (2.12), to the reduced width γ2 imposes a bound on

ΓL via eq. (2.11). Specifically, the penetration factor, PL(E, a), must be estimated to see if

the required strengths according to Figs. 2.4, 2.6 2.7, and 2.8 to solve the problem are at

all attainable. The penetration factor is given by

PL(E, a) =1

G2L(E, a) + F 2

L(E, a)(2.16)

where GL(E, a) and FL(E, a) are Coulomb wavefunctions.

We note that the Coulomb barrier penetration factor decreases as the energy of the

projectile and/or the channel radius, a, increases. For a narrow resonance, the relevant

projectile energy is E ≈ Eres, which is set by nuclear experiments (where available) and

49

their uncertainties. The channel radius corresponds to the boundary between the compound

nucleus in the resonant state and the outgoing / incoming particles. Therefore, the channel

radius depends on the properties of the compound state and the particles into which it

decays.

Consider the case of 7Be+d, which has resonance energy, Eres = 220±100 keV and initial

angular momentum, Linit = 1. A naive choice for the channel radius is the “hard-sphere”

approximation,

a12 = 1.45 (A1/31 + A

1/32 )fm (2.17)

which gives a27 = 4.6 fm. Using the Coulomb functions, Γ1 is of order a few keVs. The

corresponding strength, Γeff should be essentially the same and we further gain a factor of 6

from the spin of this state. This suggests by using figure 2.4, that this resonance should fall

short of the width required to solve or even ameliorate the problem.

However, reactions involving light nuclides including A = 7 are found to have channel

radii exceed the hard-sphere approximation Cyburt & Pospelov (2012). We thus consider

larger radii and find that for values higher than around 10 fm, we get a width which has the

potential to change the 7Li abundance noticeably. The Wigner limit

a2 =3~

2

2 µ Eres(2.18)

gives a larger radius, a27 = 13.5 fm, which gives one a better chance of solving the problem.

This is consistent with the conclusions drawn by Cyburt & Pospelov (2012).

For the 7Be + t initial state, the 18.80 MeV state of 10B has a resonance energy E =

0.131 MeV and Linit = 1. There is no experimental error bar on the resonance energy. The

hard sphere approximation gives a37 = 4.9 fm. This gives a width, Γ1, which is less than a

tenth of a keV, and is orders of magnitude lower than what is needed. In the spirit of what

we did in the earlier case, using eq. (2.18) gives a channel radius, a37 = 15 fm improving

50

the situation by almost 2 orders of magnitude in Γ1. If, in addition to increasing a37, the

resonance energy were to be higher by 100 keV, then Γ1 could be large enough to change

the 7Li abundance noticeably.

The 7Be+ 3He initial state will have 10C as the compound state. The structure of the 10C

nucleus is not well studied experimentally 20(Ajzenberg-Selove, 1990) nor theoretically. In

particular, we are unaware of any published data on 10C states near the 7Be + 3He entrance

energy, i.e., states at or near Eex(10C) ≈ Q(7Be + 3He) = 15.003 MeV. To our knowledge,

there has not been any search for narrow states in this region. The potential exit channels

of importance are 9B + p and 6Be + α. Because Jπ(3He) = 1/2+ Jπ(7Be) = 3/2−, to have

Linit = 0 and thus no entrance angular momentum barrier would require the 10C state to

have

Jπ = (1 or 2)− (2.19)

Because Jπ(9B) = 3/2−, the entrance channel spin and parity required to give Linit = 0 will

also allow Lfin = 0 for the 9B+p. On the other hand, in the final state 6Be+α both 6Be and

4He have Jπ = 0+. Thus if the putative 10C state has Jπ = 1−, this forces the 6Be + α final

state to have Lfin = 1, and thus this channel will be suppressed by an angular momentum

barrier relative to 9B + p.

Using eq. (2.17), we again get a37 = 4.9 fm. Taking Linit = 0 and E = 0.2 MeV, Γ0 is

about 10−3 keV and is extremely small. However, the penetration factor is highly sensitive

to the channel radius and a relatively small increase in a increases the width by orders of

magnitude. Increasing the energy does reduce the penetration barrier, but a higher width

is required due to thermal suppression. In order to get a sizable width, which is required to

solve the problem according to figure 2.8, a37must be >∼ 30 fm. At this energy, this radius

is somewhat larger than what is afforded by Eq. 2.18.

20http : //www.tunl.duke.edu/nucldata/figures/11figs/11 06 1990.gif.

51

Compound Nucleus, Initial Linit Lfin Eres Γtot Exit Exit ChannelJπ, Eex State Channels Width

9B, (5/2+), 16.71 MeV 7Be + d 1 0 219.9 keV unknown p+ 8Be∗

(16.63 MeV) unknown1 α + 5Li unknown

10B, 7Be + t 1 1 130.9 keV < 600 keV p+ 9Be∗

(11.81 MeV) unknown2+, 18.80 MeV 1 3He unknown

2 α unknown10C, 7Be + 3He unknown unknown unknown unknown p unknown

unknown unknown (Q = 15.003 MeV) α unknown3He (elastic) unknown

Table 2.11: This table lists surviving candidate resonances.

2.6 Discussion and Conclusions

The lithium problem was foreshadowed before precision cosmic microwave background data,

was cast in stark light by the first-year WMAP results, and has only worsened since. While

astrophysical solutions are not ruled out, they are increasingly constrained. Thus, a serious

and thorough evaluation of all possible nuclear physics aspects of primordial lithium pro-

duction is urgent in order to determine whether the lithium problem truly points to new

fundamental physics.

Reactions involving the primordial production of mass-7, and its lower-mass progenitor

nuclides, are very well studied experimentally and theoretically, and leave no room for sur-

prises at the level needed to solve the lithium problem Cyburt & Pospelov (2012); Boyd

et al. (2010b); Cyburt et al. (2003). Lithium destruction reactions are less well-determined.

While the dominant destruction channels 7Be(n, p)7Li and 7Li(p, α)α have been extensively

studied, in contrast, the subdominant destruction channels are less well-constrained.

We therefore have exhaustively cataloged possible resonant, mass-7 destruction channels.

As evidenced by the large size of Tables 2.1–2.10, the number of potentially interesting

compound states is quite large. However, it is evident that the basic conservation laws such

as angular momentum and parity coupled with the requirement of resonant reactions to be

2–3 times the 7Be(n, p)7Li rate prove to be extremely restrictive on the options for a resonant

solution to the lithium problem, and reduces the possibilities dramatically.

Given existing nuclear data, there are several choices for experimentally identified nuclear

52

resonances which come close to removing the discrepancy between the lithium WMAP+BBN

predictions and observations as tabulated in § 2.5. The 16.71 MeV level in 9B compound

nucleus, and the 18.80 MeV level in the 10B compound nucleus are two such candidates. It

is possible, however, that resonant effects have been neglected in reactions passing through

states which have been entirely missed. In all of the plots above, we have illustrated the

needed positions and strengths of such states, if they exist. One possibility involving the

compound state 10C is poorly studied experimentally, especially at higher energy states close

to the Q-value for 7Be + 3He.

Any of these resonances (or a combination) could offer a partial or complete solution

to the lithium problem, but in each case, we find that large channel radii (a > 10 fm) are

needed in order that the reaction widths are large enough. We confirm the results of Cyburt

and Pospelov Cyburt & Pospelov (2012) in this regard concerning 7Be + d, and we also find

similar channel radii are needed for 7Be + t, while larger radii are required for 7Be + 3He.

Obviously, nature need not be so kind (or mischievous!) in providing such fortuitious fine-

tuning. But given the alternative of new physics solutions to the lithium problem, it is

important that all conventional approaches be exhausted.

Thus, based on our analysis, quantum mechanics could allow resonant properties that

can remove or substantially reduce the lithium discrepancy. An experimental effort to mea-

sure the properties of these resonances, however can conclusively rule out these resonances

as solutions. If all possible resonances are measured and found to be unimportant for BBN,

this together with other recent work Boyd et al. (2010b), will remove any chance of a “nu-

clear solution” to the lithium problem, and substantially increase the possibility of a new

physics solution. Thus, regardless of the outcome, experimental probes of the states we have

highlighted will complete the firm empirical foundation of the nuclear physics of BBN and

will make a crucial contribution to our understanding of the early universe.

We are pleased to acknowledge useful and stimulating conversations with Robert Wiringa,

Livius Trache, Shalom Shlomo, Maxim Pospelov, Richard Cyburt, and Robert Charity. The

53

work of KAO was supported in part by DOE grant DE–FG02–94ER–40823 at the University

of Minnesota.

54

Chapter 3

Inverse-Compton Emission fromStar-forming Galaxies

Abstract

1 Fermi has resolved several star-forming galaxies, but the vast majority of the star-forming

universe is unresolved, and thus contributes to the extragalactic gamma ray background

(EGB). Here, we calculate the contribution from star-forming galaxies to the EGB in the

Fermi range from 100 MeV to 100 GeV, due to inverse-Compton (IC) scattering of the

interstellar photon field by cosmic-ray electrons. We first construct one-zone models for in-

dividual star-forming galaxies assuming supernovae power the acceleration of cosmic rays.

We develop templates for both normal and starburst galaxies, accounting for differences in

the cosmic-ray electron propagation and in the interstellar radiation fields. For both types of

star-forming galaxies, the same IC interactions leading to gamma rays also substantially con-

tribute to the energy loss of the high-energy cosmic-ray electrons. Consequently, a galaxy’s

IC emission is determined by the relative importance of IC losses in the cosmic-ray elec-

tron energy budget (“partial calorimetry”). We calculate the cosmological contribution of

star-forming galaxies to the EGB, using our templates and the cosmic star-formation rate

distribution. For all of our models, we find the IC EGB contribution is almost an order

of magnitude less than the peak of the emission due to cosmic-ray ion interactions (mostly

pionic pcrpism → π0 → γγ); even at the highest Fermi energies, IC is subdominant. The

flatter IC spectrum increases the high-energy signal of the pionic+IC sum, bringing it closer

to the EGB spectral index observed by Fermi. Partial calorimetry ensures that the overall

1This chapter matches the version which has been submitted to Astrophysical Journal, and posted versionon arxiv.org numbered on arXiv:1206.0770 and is co-authored with Brian Fields.

55

IC signal is relatively well-constrained, with only uncertainties in the amplitude and spectral

shape for plausible model choices. We conclude with a brief discussion on how the pionic

spectral feature and other methods can be used to measure the star-forming component of

the EGB.

3.1 Introduction

The window to the high-energy (> 30 MeV) gamma-ray cosmos has been open now for four

decades, with measurements of the diffuse emission by COS-B, (Caraveo et al., 1980; Mayer-

Hasselwander et al., 1980; Lebrun et al., 1982), OSO-3 satellite (Kraushaar et al., 1972)

followed by the second Small Astronomy Satellite (SAS-2)(Fichtel et al., 1975) and the En-

ergetic Gamma Ray Experiment Telescope (EGRET) (Hunter et al., 1997; Sreekumar et al.,

1998). SAS-2 was the first to reveal a diffuse, extra-Galactic gamma ray background (EGB).

More recently, the advent of the Fermi Gamma-Ray Space Telescope has substantially sharp-

ened our observational view of the EGB. With better energy and angular resolution and much

higher sensitivity than EGRET, Fermi has resolved many more gamma-ray point sources

and better determined the diffuse background and its energy dependence. The EGB data

are consistent with a power law of spectral index 2.41± 0.05 for energies > 100 MeV (Abdo

et al., 2010k).

The origin(s) of the EGB remains an open question. Contributions from active galaxies

(e.g., Strong et al., 1976a; Abdo et al., 2009; Stecker & Venters, 2011; Abazajian et al., 2011)

and star-forming galaxies (e.g., Strong et al., 1976b; Bignami et al., 1979; Pavlidou & Fields,

2002; Fields et al., 2010; Stecker & Venters, 2011; Makiya et al., 2011) are “guaranteed”

in the sense that these known, resolved extragalactic source classes must have unresolved

counterparts that will contribute to the EGB. Other possible EGB sources include truly

diffuse emission such as dark matter annihilation (Ando et al., 2007), interactions from

cosmic rays accelerated in structure formation shocks (e.g., Loeb & Waxman, 2000; Miniati,

56

2002), unresolved ordinary and millisecond pulsars (Faucher-Giguere & Loeb, 2010; Geringer-

Sameth & Koushiappas, 2012), and even Solar-System emission from cosmic-ray interactions

with Oort cloud bodies (Moskalenko & Porter, 2009).

One of Fermi’s major achievements has been to establish external star-forming galaxies

as a new class of gamma-ray sources. These detections give a global view of the gamma-ray

output as a result of star formation, complementary to the local gamma-ray images within

the Milky Way. Fermi has discovered γ-rays from “normal” galaxies undergoing quiescent

star formation, but only detected our very nearest neighbors, as anticipated (Pavlidou &

Fields, 2001). The Large Magellanic Cloud (LMC; Abdo et al., 2010i) has even spatially

resolved, and the SMC (Abdo et al., 2010d) and M31(Abdo et al., 2010f) have also been

detected. Other normal star-forming galaxies in the Local Group (including M33) have

not yet been seen (Abdo et al., 2010f; Lenain & Walter, 2011). Beyond the Local Group,

Fermi has detected starbursts galaxies characterised by very high star-formation rates, as

anticipated by Torres et al. (2004). Fermi has detected the starbursts M82 and NGC 253

(Abdo et al., 2010a) which had been detected at TeV energies by VERITAS (VERITAS

Collaboration et al., 2009) and HESS (Acero et al., 2009) respectively. In addition Fermi

has also detected the starbursts NGC 1068 and NGC 4945 (Nolan et al., 2012b).

The Fermi star-forming galaxies offer a qualitatively new probe of cosmic rays; they also

inform and calibrate efforts such as ours to understand the EGB contribution from the vast

bulk of the star-forming universe that remains unresolved. The LMC is the best resolved

individual system, and there the energy spectrum is consistent with pionic, while the spatial

distribution can be used to study cosmic-ray propagation (Murphy et al., 2012). More

broadly, the ensemble of all Fermi star-forming galaxies encodes information about global

cosmic-ray energetics and interaction mechanisms (Persic & Rephaeli, 2010; Lacki et al.,

2011; Persic & Rephaeli, 2012). In particular, Fermi reveals a strong correlation between

gamma-ray luminosity Lγ and supernova rate (or equivalently star-formation rate ψ). This

is expected if supernovae provide the engines of cosmic-ray acceleration. Remarkably, all

57

star-forming galaxies detected to date can be well-fit with a single power law Lγ ∝ ψ1.4±0.3

(e.g., Abdo et al., 2010f).

The main mechanism of gamma ray production in normal star-forming galaxies is an-

ticipated to be the same that dominates Milky Way diffuse gamma rays: pionic emission

pcrpism → ppπ0 → γγ, arising from interactions between cosmic-ray hadrons (ions) and inter-

stellar gas (Stecker & Venters, 2011; Abdo et al., 2009; Fields et al., 2010; Strong et al., 2010).

This mechanism is likely responsible for the non-linear relation between the luminosity of

Fermi galaxies and their star-formation rate. Namely, the observed correlation is consistent

with a picture (Pavlidou & Fields, 2001; Fields et al., 2010; Persic & Rephaeli, 2011) in

which the cosmic-ray proton flux is controlled by the supernova rate, the total number of

targets is set by the galaxy’s gas mass, and the gas and star-formation are linked by the

Schmidt-Kennicutt relation (Kennicutt, 1998).

Here, we examine the cosmological contribution of star-forming galaxies by the inverse-

Compton (IC) scattering ecrγis → eγ of cosmic-ray electrons on the interstellar radiation

field (ISRF). In order to do so, we construct a one-zone model of star-forming galaxies.

We normalize the IC emission (per unit star formation) in our model differently for normal

and for starburst galaxies. For normal galaxies we match to the Milky Way IC emission

as computed in GALPROP (Strong et al., 2010), while for starburst galaxies we set the

template matching M82 star formation rate and cosmic ray propagation consistent with

multiwavelength data just as other authors do (eg., Stecker & Venters, 2011; Lacki et al.,

2012; Paglione & Abrahams, 2012).

The IC gamma rays are produced by upscattering of interstellar radiation by high-energy

cosmic-ray electrons (Felten & Morrison, 1963; Felten, 1965; Brecher & Morrison, 1967).

Inverse-Compton scattering also represents an important energy loss mechanism for these

electrons; the other important losses are bremsstrahlung and synchrotron. The relative

importance of these losses depends on the cosmic-ray energy and on interstellar radiation

and matter densities. Where inverse Compton losses dominate, the energy injected into

58

cosmic-ray electrons is ultimately re-emitted as IC gamma-ray photons. This equality of

energy loss and energy output is known as calorimetry, and was first described in the context

of high-energy electrons by Felten (1965) and Voelk (1989) and explored in detail for diffuse

high-energy gamma rays by cosmic rays by Pohl (1994) and by Strong et al. (2010). If the

other losses are negligible then we have perfect calorimetry. But realistically, other losses

compete, and so the gamma-ray energy output is reduced by the IC fraction of the energy

losses by the cosmic rays.

In contrast to the loss-dominated propgations of cosmic-ray electrons, cosmic-ray hadrons

(mainly protons, as well as other ions) in the Milky Way suffer losses dominated by escape.

The pionic emission from normal star-forming galaxies is thus not calorimetric; however, in

the dense central regions of starburst galaxies, proton inelastic interactions can dominate

losses and lead to calorimetry (Lacki et al., 2010; H. E. S. S. Collaboration et al., 2012). In

the Milky Way, pionic emission dominates the total Galactic gamma-ray output (luminosity),

exceeding IC emission by factors of up to ∼ 5 in GALPROP calculations (Strong et al., 2010).

Our analysis will show that the inverse-Compton component from star-forming galaxies over

cosmological volumes is nearly an order of magnitude lower than the peak of the pionic

component.

The paper is organised as follows. Section 3.2 gives an order-of-magnitude estimate of the

inverse Compton EGB contribution by star-forming galaxies, and serves as an overview to our

paper. We first build a template for the IC emission from individual star-forming galaxies,

treating normal and starburst systems separately. We start with the various components of

the background interstellar photon field §3.3. These interstellar photons serve as scattering

targets for the cosmic ray electrons, whose propagation is explained in §3.4. For the highest

energy background photons and electrons, the scattering occurs in the Klein-Nishina regime,

which affects the emergent spectrum and is detailed in §3.5.1. Our one-zone model for the

inverse-Compton luminosity from a single galaxy is presented in §3.5.2. In §3.6, the total

intensity over cosmological volumes is calculated and compared with the pionic component.

59

Section 3.7 discusses the implications of our results.

3.2 Order-of-Magnitude Expectations

An order-of-magnitude estimate of our final result will help frame key physical issues and

astrophysical inputs. Our goal is to find the EGB contribution from inverse Compton scat-

tering in star-forming galaxies throughout the observable universe. We thus calcualte the

gamma-ray specific intensity, IE , at energy E. For photons up to at least <∼ 30 GeV, the

universe is optically thin;2 thus, the intensity is simply given by the line-of-sight integral

IE ≈ c

los

dℓ Lγ ≈ LγdH

4π(3.1)

where Lγ is the luminosity density or cosmic volume emissivity of the galaxies, and dH =

c/H0 = 3000h−1 Mpc is the Hubble length. The cosmic luminosity density of a distribution of

IC-emitting, star-forming galaxies can be expressed as a product of the luminosity function,

dngal/dLγ and the luminosity of each individual galaxy, Lγ:

Lγ =

Lγdngal

dLγ

(Lγ) dLγ = 〈nLγ〉 ≈ ngalLE (3.2)

where LE is the luminosity of an average galaxy at energy E.

Within a single galaxy, the gamma-ray luminosity is an integral

Lγ(Eem) =

qic dVe (3.3)

over the volume in which cosmic-ray electrons propagate. Here the inverse-Compton volume

emissivity qic depends on the product of the targets and projectile densities, and the inter-

action cross section. The density of targets is simply the number density of the interstellar

2We negelect the interactions with the extragalactic background light, leaing to pair production γγ →e+e−; these effects could be important above ∼ 100 GeV (e.g., Salamon & Stecker, 1998; Abdo et al., 2010e;Stecker et al., 2012).

60

photons, nisrf . The projectiles are cosmic-ray electrons, with flux density φe. The cross sec-

tion is that for inverse Compton scattering, dσic/dEem. Therefore, the emissivity for a single

galaxy is given as,

qic =

φedσic

dEem

nisrf

(3.4)

with the angle brackets indicating averaging over the cosmic-ray and background photon

energies.

We will assume that cosmic-ray acceleration is powered by supernova explosions. In-

deed, the cosmic-ray/supernova link historically has been better established for cosmic ray

electrons than for hadrons, thanks to radio (e.g., Webber et al., 1980; Strong et al., 2011;

Bringmann et al., 2012) and X-ray synchrotron (Uchiyama et al., 2007; Helder et al., 2009,

e.g.,) measurements. More recently Fermi-LAT (Volk et al., 1996; Ahlers et al., 2009) mea-

surements of pionic gamma rays provided the most direct evidence for supernova acceleration

of protons and other ions. Consequently, a galaxy’s cosmic-ray injection rate is proportional

to its supernova rate, qcr ∝ RSN. And because the supernova rate itself traces the galactic

star-formation rate ψ, we conclude that qcr ∝ ψ.

Cosmic-ray electron propagation is dominated by energy losses in the form of inverse

Compton, bremsstrahlung and synchrotron processes. Each contributes to a total energy

loss rate btot = |dEe/dt|. The IC loss rate is proportional to the background photon energy

density: bic ∝ Uisrf ∝ nisrf . The cosmic-ray flux φe is inversely proportional to the total energy

loss rate, btot φe ∝ (qe/ btot). Thus, IC emissivity, qic is lower than rate qe of energy injection

into cosmic-ray electrons, by the fraction bic/btot of energy losses in IC, qic ∝ (bic/ btot)qe. This

is the case of partial or fractional calorimetry where qic/qe ∝ bic/btot: IC photons trace the

portion of cosmic-ray energy lost via this mechanism (Pohl, 1994). Even the approximate

validity of calorimetry is sufficient that the IC gamma ray luminosity is a fairly robust

calculation.

61

Due to partial calorimetry, the IC volume emissivity is qic ∝ qe, and integration over all

of the supernovae acting as cosmic-ray accelerators gives the IC luminosity LE ∝ RSN ∝ ψ.

Physically, a fixed fraction of each supernova’s energy goes into cosmic-ray electrons and

ultimately into IC photons.

For normal star-forming galaxies, this scaling can then be compared to GALPROP esti-

mates for the total Milky-Way IC luminosity (Strong et al., 2010), which determines the IC

output per unit star-formation. We can then find the luminosity for any star-forming galaxy

at a fiducial energy of E = 1 GeV as,

E2 LE = E2 LE,MW

(

ψ

ψMW

)

≈ 1040 GeV2 sec−1 GeV−1

(

ψ

1 M⊙ yr−1

)

(3.5)

where ψMW is the Milky-Way star-formation rate.

From eqns. (3.1) (3.2), and (3.5), we estimate the IC contribution to the EGB intensity

at 1 GeV as

E2 IE∣

ic,1GeV≈ LE,MW ngal dH

4 π

(

ψ

ψMW

)

≈ LE,MW dH

4 π ψMW

ρ⋆(1)

≈ 3 × 10−8 GeV cm2 sec−1 sr−1 (3.6)

where ρ⋆(1) = ψngal ≈ 0.1 M⊙ yr−1 Mpc−3 is the cosmic star formation rate, evaluated at

z = 1, near its peak (Horiuchi et al., 2009a). For comparison, the pionic model of Fields

et al. (2010) gives an intensity E2 IE|π→γγ,1GeV ∼ 3×10−7 GeV cm2 sec−1 sr−1, with a factor

∼ 2 uncertainty in normalisation. The EGB measured by Fermi-LAT is E2 IE|obs,1GeV = 6×

10−7 GeV cm2 sec−1 sr−1 (Abdo et al., 2010k). Thus, the pionic component alone dominates

the overall amplitude from the star-forming galaxies, which is an important contribution to

the total observed flux. Not evident from our estimate is that the shape of the star-forming

spectrum improves upon addition of the IC component, as we will see in §3.6.

62

3.3 Targets: Interstellar Photon Fields

Figure 3.1: The interstellar photon field (ISRF) energy density, shown as a function of thewavelength for normal (left panel) and starburst galaxies (right panel). (a) Left: Milky-Waytemplate for normal galaxies The model includes thermal components for starlight (optical),dust (IR) and the CMB as in Cirelli & Panci (2009). It includes another thermal componentfor the starlight in the UV. The black (fiducial), red and blue curves represent the curves for(R, h) = (0,5 kpc), (4 kpc,0) and (0,0) respectively. (b) Right: M82 template for starburstgalaxies. The solid black curve is a model (de Cea del Pozo et al., 2009) for a starburst ISRFdominated by IR emission due to the dusty nuclei. It has two thermal components, a coolone at 45 K and a warm one at 200 K as in de Cea del Pozo et al. (2009). These correspondto IR luminosity of 1010.5Lsol and a core radius of 200 pc. It is compared with the fiducialISRF of the Milky Way type normal galaxy shown in dashed black.

Cosmic-ray electrons in a galaxy will inverse-Compton upscatter ambient photons, which

are commonly called the interstellar radiation field (ISRF). Each galaxy’s ISRF is a rich

function of both energy and geometry. We simplify this complex reality by treating a galaxy

as a single zone for the purposes of cosmic-ray propagation and gamma-ray emission. Thus,

our ISRF is meant to give a sort of effective volume average of the radiation field encountered

by cosmic-ray electrons. A galaxy’s ISRF will evolve with redshift, and so for our single-

galaxy templates we must also create a prescription for how the different ISRF component

evolve with z.

63

Our choice of ISRF for normal galaxies follows those of models for the Milky Way, which

we adjust at other cosmic epochs via scaling arguments. These scalings use the cosmic

star-formation rate ρ⋆(z), whose change with redshift we characterize by the dimensionless

function

S(z) ≡ ρ⋆(z)

ρ⋆(0)(3.7)

where today S(0) = 1.

For the ISRF at the present day (redshift z = 0), we adopt but extend the model proposed

by Cirelli & Panci (2009), which consists of three thermal components corresponding to the

infra-red (or dust), optical, and the Cosmic Microwave Background (CMB). To this we add a

fourth thermal component for the UV. Our goal is to build a simple model that can reproduce

the inverse Compton luminosity of the Milky Way according to Strong et al. (2010); their

calculation is based on a far more detailed, spatially-dependent ISRF as implemented in the

GALPROP code (Strong et al., 2000; Porter et al., 2008). The strengths of the UV, IR and

the optical relative to the CMB decreases with Galactocentric radial distance R and also

with height h above the Galactic plane. This is seen in Fig. 3.1.

We may express the ISRF specific energy density duisrf/dǫ = ǫ dnisrf/dǫ as

λduisrf

dλ= ǫ

duisrf

dǫ=

4∑

i=1

fi1

π2

ǫ4

exp(ǫ/Ti) − 1(3.8)

a sum of Planck terms, with dimensionless weights fi for the different background compo-

nents: i = UV, optical, IR and the CMB. The UV and optical component are from starlight,

and the IR comes from starlight reprocessed by dust. Our model, as seen in the Fig. 3.1

thus includes the effect of the major components in the GALPROP ISRF, which dominate

the energy density. For the IR, we tried a non-thermal model that includes effect of both

scattering and absorption by dust particles, but the effects on our gamma ray luminosity

model are insignificant. Therefore, we retain the thermal model for IR. We also verified that

including the optical-IR transition where the non-thermal lines appear has negligible effect

64

on the final gamma ray luminosity.

Figure 3.1 shows the ISRF at different locations in the Galaxy. The optical and dust

components are taken from Cirelli & Panci (2009). We add another thermal component for

the UV. The black, red and blue curves correspond to the positions (R, h) = (0, 5), (4,0) and

(0,0) kpc. The peaks due to starlight and IR change with position in the Galaxy, while the

CMB contribution peaked at λ ≈ 1000 microns remains the same. At the Galactic centre, the

energy densities of these latter components are greater than the CMB, with optical photons

dominating. Above the Galactic plane, the CMB becomes comparable in energy density and

eventually dominates at very large h.

Table 3.1 lists the temperatures of the various ISRF components along with their relative

weights at zero redshift, fi(0) for the normal galaxies. We chose as a fiducial model the rela-

tive strengths for regions corresponding to 10 < |b| < 20 and 0 < l < 360, corresponding

to the ‘10-20’ model of Cirelli & Panci (2009). This is for the position (R, h) = (0, 5) Kpc

represented by the black curve in Fig. 3.1. The UV component at this position is scaled

from that in the Galactic Centre, (Porter et al., 2008) in the same proportion as the optical

component in these positions. This choice of model, with the overall normalisation adjusted,

reproduces the Milky Way IC luminosity as a function of energy from Strong et al. (2010).

The ISRF in starburst nuclei is dominated by the IR peak, which renders other compo-

nents relatively unimportant. Our one zone ISRF template is based on the two-component

model of de Cea del Pozo et al. (2009). The starburst ISRF is dominated by a thermal

peak corresponding to a temperature, T = 45 K, according to de Cea del Pozo et al. (2009).

The energy density is related to the IR luminosity, L from the nucleus of radius R as,

Uisrf = L/2πR2c (Lacki et al., 2010). Picking a normalisation based on the energy density

Uisrf ≈ 1100 eV cm−3, set by luminosity, L = 1010.5L⊙, and radius, R = 200 pc, we get the

green curve in the right panel of Fig. 3.1. However, according to de Cea del Pozo et al.

(2009), the dust emission fits better with an additional warm component with temperature

of 200 K. This is our fiducial model. The choice of the fiducial ISRF is displayed in solid

65

black in the right panel of Fig. 3.1. It is also summarised in the table 3.1. And we pick the

overall normalisations fi(0) in order match the same energy density as for the one component

case.

Table 3.1: ISRF Parameters for normal galaxiesMilky Way UV Optical IR CMB

Fiducial : fi(z = 0) 8.4 × 10−17 8.9 × 10−13 1.3 × 10−5 1Galactic Centre fi(z = 0) 1.6 × 10−15 1.7 × 10−11 7 × 10−5 1

T (z = 0) [K] 1.8 × 104 3.5 × 103 41 2.73

Table 3.2: ISRF Parameters for starburst galaxiesStarburst M82 type UV Optical IR (cool) IR (warm) CMBFiducial : fi(z = 0) 3.2 × 10−15 0.0 4.22 × 10−2 3.61 × 10−5 1

T (z = 0) [K] 1.8 × 104 3.5 × 103 45 200 2.73

Because we are interested in inverse Compton emission from star-forming galaxies over

all of cosmic history, we must specify the cosmic evolution of the ISRF. This will depend on

which stars contribute to starlight and dust scattered light. The detailed present-day Milky-

Way model in Strong et al. (2000) uses the stellar luminosity functions from Wainscoat et al.

(1992). The UV and IR components derive from short-lived massive stars, and the resulting

starlight density will be proportional to the star-formation rate. Thus, we adopt the scalings

f2(z) = fUV(z) = fUV(0) S(z),

f3(z) = fIR(z) = fIR(0) S(z) (3.9)

The optical component is less trivial and evolves differently for different stellar popula-

tions. Massive supernova progenitors, as well as AGB progenitors, are short-lived compared

to >∼ 1Gyr timescales for cosmic star formation. Consequently, the optical radiation density

of these stars would scale as

f2(z) = fop(z) = fop(0) S(z) , (3.10)

in step with the UV and IR scaling. For reasonable initial mass functions and star-formation

66

histories, these stars should dominate the optical ISRF. But lower-mass main sequence stars

and their red giant descendants have cosmologically long lifetimes; this means that their

starlight density at any epoch would scale as the integrated star-formation rate. Therefore,

their contribution to the optical component would scale as

f ′op(z) = fop(0)

∫∞

zdz| dt

dz| ρ⋆

∫∞

0dz| dt

dz| ρ⋆

, (3.11)

This would give a lower optical component and decrease the IC intensity as we will see. We

will adopt eq. (3.10) for our fiducial model, but we will explore the effects of using eq. (3.11);

these cases will bracket the true evolution.

The redshift dependence of the CMB is precisely known and can be expressed entirely in

terms of its temperature,

TCMB(z) = (1 + z) TCMB(0) (3.12)

with f4(0) = fCMB(0) = 1. The CMB energy density thus grows rapidly as Ucmb ∝ (1 + z)4.

3.4 Projectiles: Cosmic-Ray Electron Source and

Propagation

An electron with energy Ee = γmc2 scattering off a background photon of energy ǫ yields an

IC photo with energy Eγ ∼ γ2ǫ. In order to produce gamma rays around the Fermi range

Eγ ∼ 100 MeV to ∼ 300 GeV , electrons with energies of few GeV to few TeV are required,

depending on the energy of initial un-scattered photon. These are the electrons of interest

to us.

High-energy cosmic-ray electrons obey a complex transport equation (e.g., Strong &

Moskalenko, 1998). This in general would include effects of diffusion, convection, escape,

radiative losses, etc. However, our approach is to create the simplest one-zone model that

67

captures electron source spectrum and their most important losses.

We model cosmic-ray electron injection for normal galaxies via the source spectrum from

the Plain Diffusion (PD) model of Strong et al. (2010). In this model, electrons are accel-

erated with a broken power-law injection spectrum having qe(Ee) ∝ E−1.8e up to a break

at Ee,break = 4 GeV, and above this qe(Ee) ∝ E−2.25e . Electrons below Ee,break can produce

inverse Compton gamma rays at the low end of the Fermi range, when upscattering optical

and UV photons. Gamma rays at higher energies are produced by electrons above the break

energy. Thus, it is important to include the injection break in our calculations. Including

a cutoff at 2 TeV according to Strong et al. (2011), results in a rollover at high energies

improving the agreement with the GALPROP model beyond 100 GeV.

For starburst galaxies the injected cosmic-ray electrons important for IC emission are

difficult to constrain independently from gamma-ray emission, and thus are more uncertain.

For starburst galaxies, we use M82 as a template, and thus scale the injected spectrum up

from that of the Milky Way by the ratio ψM82/ψMW of star-formation rates, using ψM82 =

ψ8 M⊙ yr−1 in accordance with the total IR luminosity Sanders et al. (2003) as used by

Lacki et al. (2012). For our fiducial model we adpot an injection spectrum that is a single

power law qe(Ee) ∝ E−2.1e , although other indices ranging from 2.0 to 2.4 are permissible

(Paglione & Abrahams, 2012). The high-energy cutoff being less constrained (Lacki et al.,

2012) we pick fiducial value of 2 TeV, same as the Milky Way. Our fiducial model is directly

comparable with Paglione & Abrahams (2012).

Milky-Way cosmic-ray electrons experience an interstellar diffusion coefficient, D0 ∼

1028cm2sec−1 (Strong et al., 2010), so the diffusion length for electrons over the Compton

loss time-scales is ℓdiff ∼ 0.01 − 1 kpc, depending on the energy. We thus see that electrons

will not venture far from the Galaxy, but can range vertically outside the Galactic midplane.

Hence, very few electrons will be lost to escape, in strong contrast to ions for which escape

dominates (Pohl, 1994). We thus will ignore electron escape in our analysis.

However, since the electrons do propagate throughout and above the stellar and gaseous

68

disk, the electron population sees a wide variety of radiation fields. Our single choice is a

crude approximation meant to represent a sort of average ISRF. As a result, the electron

propagation equation takes the “thick target” form

∂tφe(Ee) = qe(Ee) +

∂Ee

[b(Ee)φe(Ee)] (3.13)

where qe(Ee) is the source term and b(Ee) = |dEe/dt| is the total rate of energy loss. The

equilibrium (i.e., ∂tφe = 0 steady state) solution of the cosmic ray flux is

φe(Ee) =1

b(Ee)

Ee

dE ′e qe(E

′e) =

qe(> Ee)

b(Ee)(3.14)

As usual in the thick-target limit, the electron flux is directly proportional to the energy-

integrated injection rate qe(> Ee), and inversely proportional to the total energy loss rate.

In general, the total energy loss rate of the electrons is given by sum of the inverse Compton,

synchrotron, bremsstrahlung and ionisation losses,

btot(Ee) = bic(Ee) + bsync(Ee) + bbrem(Ee) ; (3.15)

we now address each in turn. The ionisation loss is unimportant throughout the entire

electron energy range of interest here.

The inverse Compton loss in the Thomson regime takes the simple form bic(Ee)Thoms

=

43σT c Uisrf

(

Ee

m c2

)2. However, the full Compton cross section takes the Klein-Nishina form

described in more detail in §3.5.1. Corrections to the Thomson limit become important

when ∆ǫ = 4 ǫγ/m c2 ≫ 1; in a starlight-dominated ISRF with ǫ ∼ 1 eV, this occurs for

electrons with γ ≫ 105. This energy regime is important for our calculation and therefore,

we calculate the inverse Compton loss rate by combining the Klein-Nishina cross section

69

self-consistently with the (redshift-dependent) ISRF:

bic(Ee) =

∫ ∞

0

dǫdnisrf

dǫ(ǫ)

dEem (Eem − ǫ)dσic

dEem

= 3 σT c

∫ ∞

0

dǫdnisrf

dǫ(ǫ) ǫ

∫ 1

(m c2/2Ee)2dq

(4(

Ee

m c2

)2 − ∆ǫ)q − 1

(1 + ∆ǫq)3

×(

2q ln(q) + (1 + 2q)(1 − q) +1

2

(∆ǫq)2

1 + ∆ǫq(1 − q)

)

(3.16)

as in Cirelli & Panci (2009), where

q =Eem

4ǫγ2(1 − Eem/γm c2). (3.17)

Note also that for the highest-energy electrons, inverse Compton losses become catastrophic,

i.e., in each scattering event the fractional changes in the electron energy approach δEe/Ee ∼

1 (e.g., Blumenthal & Gould, 1970). In this case the losses can no longer be treated as a

smooth function bic(Ee) that averages over many scatterings each with δEe/Ee ≪ 1; we do

not include these effects, which impact gamma-ray energies of Eem >∼ 1 TeV.

The relative importance of the losses determines the propagated electron spectrum (Fel-

ten, 1965; Pohl, 1993). For our fiducial ISRF model, the bremsstrahlung, synchrotron and

Compton losses are all important, depending on the part of the spectrum. The rates are

given in the Appendix B.1. For electron energies lower than ∼ 1 GeV, bremsstrahlung is

dominant scaling as bbrem ∝ Ee, with Compton losses being comparable close to the break

energy at Ee = 4 GeV, scaling as bic ∝ E2e . Since, the electron injection spectral break and

the bremsstrahlung–Compton transition are fairly close, instead of seeing two breaks corre-

sponding to these two cosmic ray propagation regimes, only one is seen near the break energy

as described in §3.4. For large Ee, inverse Compton losses drop off to a logarithmic energy

dependence due to Klein-Nishina suppression, while synchrotron losses maintain bsync ∝ E2e .

Thus synchrotron losses dominate over Compton at TeV and higher energies.

The losses themselves evolve with redshift, through the dependences on interstellar den-

70

sities, i.e., the ISRF energy density Uisrf , the interstellar magnetic energy density Umag =

B2/8π, and the number densities nism of interstellar particles. The magnetic field couples

to all cosmic rays, for which the ions dominate the energy density. This coupling is likely

responsible for the approximate equipartition of magnetic field and cosmic-ray energy den-

sities observed in the local interstellar medium. We thus assume that the magnetic field

energy density scales with the cosmic-ray ion flux, and in turn with the star-formation rate:

Umag ∝ S(z). For our fiducial normal galaxy model, we use B0 = 4 µ G, to match the model

of Strong et al. (2010), which is comparable with the typical value in Strong et al. (2011).

We take interstellar particle densities, which control bremsstrahlung losses, to scale as

ni(z) = ni,0(1 + z)3 (3.18)

This can be viewed a consequence of the disk radius scaling with the cosmic scale factor

Rdisk ∝ a ∝ (1 + z)−1 in the Fields et al. (2010) model, and to reflect proportionality

between galaxy disk and dark halo sizes. The fiducial number densities of neutral and ionised

hydrogen, i.e. nH II,0 = nH I,0 = 0.06 cm−3. The helium content of ionised and neutral gas

is also included, with yHe = 0.08. These parameters are broadly consistent with values at

intermediate heights h above the Galactic plane, as befits the volume average represented

by our one-zone model. For reference, the values of nism and B used at the Galactic centre

are much higher at B0 = 8.3 µG and nH II,0 = nH I,0 = 0.12 cm−3.

For starburst galaxies, the magnetic fields and the interstellar densities are higher. The

uncertainty attached with them are also higher compared to Milky Way type normal galaxies.

The fiducial values for magnetic field chosen here is B = 300 µG, although different values

are possible depending on the radio observations of electrons and assumptions regarding

the e/p injection ratio (Lacki et al., 2010; Paglione & Abrahams, 2012; Lacki et al., 2012).

Changing B will change the importance of the IC relative to synchrotron and also pionic

losses. The fiducial ISM density is nism = 600 cm−3 is comparable to the mean densities

71

from observations of M82 (Weiß et al., 2001). These parameter values are in similar range as

used by other groups such as Lacki et al. (2012); Paglione & Abrahams (2012). The redshift

dependence of these parameters is chosen to be the same as for normal galaxies.

3.5 Inverse Compton Emission from Individual

Star-Forming Galaxies

As seen in eq. (3.2), we first compute the luminosity of single galaxy and then average suitably

over the luminosity function to get the luminosity density or cosmic volume emissivity. In

this section, we address the former, focussing first on the IC emissivity within the volume of

a single galaxy, then using this to compute the total luminosity from that galaxy.

3.5.1 Inverse Compton Emissivity of a Star-Forming Galaxy

The specific IC volume emissivity within a galaxy is the rate of producing gamma ray

photons, dΓe−γ→e−γ/dEem per background photon, multiplied by the ISRF number density,

nisrf

dqγdEem

=dNγ

dV dEemdt=

dǫdΓe−γ→e−γ

dEem

dnisrf

dǫ. (3.19)

Here, the gamma ray rate per unit ISRF photon is a product of the cosmic ray flux and the

IC cross section

dΓe−γ→e−γ

dEem=dσe−γ→e−γ

dEemφe . (3.20)

that in general takes the Klein-Nishina form. The full Klein-Nishina cross section is a com-

plicated function of the energies involved. Moreover, the inverse Compton process produces

anisotropic emission if either the cosmic-ray or ISRF populations depart from isotropy. While

72

the cosmic-ray electrons of interest are well-approximated as isotropic, in a real galaxy the

photon field is certainly anisotropic as well as spatially-varying. These effects are included

in GALPROP (Moskalenko & Strong, 2000), but we will ignore them in our simple one-zone

galaxy model. In general, an anisotropic ISRF leads to an overall anisotropy in IC 3. From

figures 5(a) and 6(a) of Moskalenko & Strong (2000), the ISRF anisotropy is a ∼ 20% effect

in the regions of strong IC emission, i.e., within the diffusion length of a few kpc of the disk.

Thus, this effect will not dominate our final error budget.

For the electron spectra we consider, the Thomson limit is a good approximation for

CMB and IR photons, but fails for UV and some optical photons. Hence, we use the Klein-

Nishina result via the prescription of Jones (1968) and Blumenthal & Gould (1970). This

differential cross section is expressed as,

dEem(ǫ, Ee, q) =

3

4

σT

ǫ γ2

(

2q ln(q) + (1 + 2q)(1 − q) +1

2

(∆ǫq)2

1 + ∆ǫq(1 − q)

)

where the electron Lorentz factor γ = Ee/me and σT is the total Thomson scattering cross

section and q appears in eq. (3.17). In the Thomson limit, the last term in the above equation

become negligible.

The inverse Compton volume emissivity within a one-zone galaxy is thus

dqγdEem

=

dǫdΓe−γ→e−γ

dEem

dnisrf

dǫ=

dǫdnisrf

dEeφe(Ee)dσe−γ→e−γ

dEem(3.21)

We evaluate the emissivity numerically, by using the ISRF, cosmic ray flux density, and the

cross section mentioned above, §3.3 and §3.4. That ISRF density nisrf is the same as in the

inverse Compton energy loss equation eq. (3.19), so that we calculate the two self-consistently.

3In principle, therefore, a galaxy’s IC emission is anisotropic and depends on the galaxy orientation! Ofcourse, the EGB signal will average over many galaxies and thus over all orientations.

73

It is suggestive to recast the emissivity in eq. (3.21) in the form

dqγdEem

= Uisrf

dǫdnisrf

dǫǫ(

dΓ(ǫ,Eem)dEem

)

dǫdnisrf

dǫǫ

≡ Uisrf

1

ǫ

dΓ(ǫ, Eem)

dEem

= Uisrf

1

ǫ

qe(> Ee)

b

dEem

(3.22)

We see that the emissivity can be understood as an effective average over the ISRF en-

ergy density distribution. Here, the differential and total ISRF energy densities are uisrf =

ǫ dnisrf/dǫ and Uisrf =∫

dǫ uisrf respectively. We see that the amplitude of the gamma-ray

emissivity, qγ scales as

qγ ∼ Uisrf

bqe ∼

bicbqe

and in the Compton dominated regime simply as qe and is independent of the ISRF energy

density, Uisrf . This is a statement of calorimetry. Of course, the gamma-ray spectral shape

does remain sensitive to the ISRF even in the case of perfect calorimetry.

In reality, synchrotron and bremsstrahlung compete with inverse Compton losses at dif-

ferent energies, as seen in §3.4. In the Thomson regime, the ratio of synchrotron to inverse-

Compton losses

bsync

bic

Thomps−→ UB

Uisrf

= 0.4

(

1.1 eV cm−3

Uisrf

)(

B

4µG

)2

, (3.23)

is energy-independent; here the numerical factors are for normal galaxies. For our fidu-

cial magnetic field value, this ratio is not far from unity and hence, the shape does not

change much for different parameters describing the ISRF and the magnetic field. On

the other hand, for Ee >∼ 1 TeV, the IC losses are in the Klein-Nishina limit and grow

only logarithmically in energy; then the synchrotron losses dominate In this case, we have

qγ ∼ (Uisrf/Umag)qe. Thus, for high-energy cosmic-ray electrons, the calorimetric approxima-

tion increasingly breaks down.

For starbursts, the syncrotron-to-IC ratio (eq. 3.23) is bsync/bic ≃ 6 in our fiducial model.

Thus, synchrotron losses dominate IC for all electron energies for such high magnetic fields.

This decrease in emissivity due to synchrotron losses is partially compensated by the in-

74

creased star formation rate in starbursts.

The competition among losses depends on the way the interstellar densities evolve with

redshift. Short-lived, massive stars dominate most of the ISRF except possibly for some of

the optical range, and so the energy density scales with the star-formation rate Uisrf ∝ S(z).

Due to cosmic-ray ion equipartition, we take Umag ∝ S(z) as well, and thus the Uisrf/Umag

ratio doesn’t change dramatically throughout the cosmic history. Similarly, we also include a

strong evolution nism ∝ (1+ z)3 evolution to the interstellar particle density, which increases

rapidly towards z ∼ 1, roughly in step with the cosmic star-formation rate. Thus, the

inverse-Compton/bremsstrahlung ratio of energy losses remains roughly constant as well out

to z ∼ 1. The net result is that the spectral shape of inverse Compton emission does not

evolve dramatically with redshift in our model.

3.5.2 Total Inverse-Compton Luminosity from a Single Galaxy

As seen in §3.2, knowing the specific inverse Compton emissivity dqγ/dEem within a single

star-forming galaxy, we can calculate the specific IC luminosity of the galaxy. Namely,

Lγ(Eem) =dNγ,ic

dt dEem=

dqγdEem

dVISM (3.24)

and we have seen that qγ ∝ (bic/btot)qe. Following Fields et al. (2010), we take supernovae

as the engines of cosmic-ray acceleration. Thus we scale the electron injection rate with

the supernova rate, i.e., we have Lγ ∼∫

qe dVISM ∝ RSN ∝ ψ. This linear proportionality

captures the main dependence of IC emission on star-formation rate.

The single-galaxy IC luminosity takes the separable form

Lγ(ψ, z) =ψ

ψgalLgal(z) (3.25)

where we have suppressed the energy dependence for clarity. Here we explicity display the

dominant sensitivity to star-formation, via the overall linear proportionality to the star-

75

Low B

Low electron cutoff

High injection index

Figure 3.2: The inverse-Compton gamma-ray luminosity spectrum for our single-galaxytemplates. (a) Left: Milky-Way template for normal galaxies. The solid black curve repre-sents our fiducial model, with an ISM density nH I,0 = nH II,0 = 0.06 cm−3, magnetic fieldB0 = 4 µG with a cutoff in the cosmic ray spectrum at 2 TeV. This is for the ISRF consistentwith (R, h) = (0, 5) kpc. This model is designed to provide a good fit to the GALPROP PlainDiffusion model of Strong et al. (2010, dot-dashed blue curve). The solid red curve showsthe effect of reducing B to 2 µ G. The dotted red curve is for nH I,0 = nH II,0 = 0.03 cm−3.The red dashed is includes the effect of non-thermal lines in the ISRF. The solid blue showsthe single galaxy template for the Galactic Centre, with nH I,0 = nH II,0 = 0.12 cm−3, andB0 = 8.3 µG. Comparison of the Diffusive reacceleration case with GALPROP with ourone-zone template that fits it is shown in green. The black dotted curve shows effect ofthe modifying the optical component of the ISRF to include effect of finite lifetime of stars.(b) Right: M82 template for starburst galaxies. The fiducial curve for M82 type starburstis shown in black with nH I,0 = nH II,0 = 300 cm−3, B = 300 µG, and a cutoff at 2 TeV.And the ISRF is shown by the thick black curve in Fig. 3.1. And the magenta curve herecorresponds to the ISRF with higher IR temperature also shown in magenta. Variations inB, n, and spectral index are shown by the blue, green and red curves respectively.

76

formation rate. The Milky-Way model LMW(z) is for a galaxy with the present-day Milky-

Way star-formation rate, but allowing for redshift evolution of the interstellar matter and

radiation densities. Starbursts follow similar scalings, but are tied to M82.

Figure 3.2(a) shows our IC luminosity models for z = 0 for normal galaxies, along

with that of Strong et al. (2010). Our fiducial model described in §3.3, §3.4 and §3.5.1

is represented by the solid black curve. The main feature of the inverse Compton E2emLMW

spectrum is that it has a broad maximum corresponding to the break in the electron injection

spectrum at Ee,break = 4 GeV. In our the fiducial model, the optical component dominates

the ISRF, and thus we expect the injection peak to correspond to an IC peak at Eem ≈

γ2ǫ ≈ 100 MeV. This is close to what we find in our fiducial model, and is consistent with

the GALPROP results.

Our fiducial Milky Way model does a good job of reproducing the IC results of Strong

et al. (2010), over most of the Fermi energy range. Thus, our results are equivalent to

normalizing to the GALPROP luminosity for normal galaxies with the Milky Way’s star

formation rate. A fitting function for our fiducial IC model appears below in Table 3.3; the

fit is good to better than 2% over the Fermi energy range. From this fit we see that at 1

GeV the logarithmic slope is around −0.1, so that LMW ∼ E−2.1em . This is considerably flatter

than both the pionic spectrum and the observed EGB at these energies.

Fig. 3.2(a) presents other curves that illustrate the sensitivity of our model to the input

parameters. The resulting trends can be understood in terms of the competition among

the different electron energy losses in the electron propagation through to the resulting IC

spectrum. For example, decreasing the gas density suppresses the bremsstrahlung losses

relative to the total. This increases the ratio of bic/bbrem, which governs the IC spectral

shape at lower energies. The result is a boost in the IC output at lower energies, which also

shifts the IC peak towards lower energies. At high gamma-ray energies beyond ∼ few GeV,

synchrotron losses dominate, and thus the ratio bic/bsync controls the spectrum; thus we find

that reducing the magnetic field increases the luminosity.

77

The choice of ISRF is also important. Our fiducial model adopts an ISRF consistent

with the GALPROP (R, h) = (0, 5) kpc emission, as it best reproduces the IC luminosity

of (Strong et al., 2010). If we change the ISRF to that of the Galactic centre, the IC signal

is amplified, and the peak shifts to lower energies, as shown by the solid blue curve in

Fig. 3.2(a). At the Galactic centre, the magnetic energy density is increased by a factor of

4 relative to the fiducial case, but the photon energy density goes up by more than a factor

of 10. So both the fractions, bic/b and bsync/b are higher, but so is bic/bsync. And so the peak

occurs at a lower energy, but has a greater amplitude. We included the effect of non-thermal

lines in the ISRF, using the GALPROP spectral form. We find that there is only a minor

effect on the galactic luminosity as shown by the red dashed curve in Fig. 3.2(a).

Finally, we show the effect of different redshift evolution of the optical ISRF component.

Namely, the blue dotted curve in Fig. 3.2(a) shows the results if the optical component is

dominated by long-lived stars. In this model the luminosity scales with stellar mass rather

than star-formation rate: Lγ ∝ Lbg ∝M⋆. Compared to our fiducial case, the ISRF shape is

still similar (same red stars), but the redshift history of Lγ(z) ∝ M⋆(z) is different because

ρ⋆(z) builds up more slowly than ρ⋆(z). We see from Fig. 3.2(a) that fixing our single galaxy

template to z = 1 at the peak of star formation activity, reduces the galactic luminosity due

to lower stellar mass.

Thus, there are several uncertainties in our simple model of the gamma ray spectrum for

a single, Milky-Way like galaxy, such as choice of the average ISRF, the electron density, and

the interstellar magnetic field. However, for reasonable interstellar models, these all lead to

modest changes in the normalisation of the luminosity in the Fermi energy range, at most

tens of percent.

Larger changes would be possible if one were to depart from the single galaxy template

in Strong et al. (2010), which we try to reproduce and that is based on a wealth of Milky-

Way data. Allowing oneself sufficient freedom, the inverse Compton EGB signal can be

adjusted by changes, e.g., in the Uisrf/UB and Uisrf/nism ratios that control the degree of

78

electron calorimetry. Of course, this would then drive the system away from the rough

energy equipartition observed in the Milky Way. Even then, large increases in these ratios

would only increase the completeness of electron calorimetry and would raise the IC signal

by a factor <∼ 2; on the other hand, large decreases in these ratios would spoil calorimetry

and could substantially lower the IC luminosity. Thus, the cosmological prediction of the IC

contribution of the star-forming galaxies to the EGB should be fairly robust, unless there

are departures from the equipartition implicit in the Milky-Way-based normalization.

We turn now to the template for starbursts, which we based our models of M82. The

starburst GeV gamma-ray emission shown Fig. 3.2(b) is characterised by the lack of a peak,

because the underlying electron spectrum is a single power law lacking a break. The overall

normalisation of the fiducial model shown by the black curve is essentially a result of the

contest between the higher star formation rate and the lower fraction of IC losses as expressed

by,

Lnormal

LSBG≈

(

ψMW

ψM82

)(

Uisrf,MW

Uisrf,M82

)(

Uisrf,M82 + UB,M82

Uisrf,MW + UB,MW

)

(3.26)

, assuming the acceleration e/p ratio is the same for normals and starbursts. Equation 3.26

predicts that an M82-type starburst show have an IC luminosity that is a factor ∼ 3.6 larger

than our Milky-Way template, which is in agreement with the results show in Fig. 3.2(a)

and (b). Moreover, the differences in the Fig. 3.2(b) curves for different M82 parameters can

be understood in the same way as the same trends in B, nism and ISRF discussed above for

the Milky-Way template.

The M82 template emission at GeV energies is a factor of few higher compared to the

Milky Way template. However, the M82 emission is proportional to its star-formation rate,

which is higher than that of the Milky Way by a factor ∼ 8. Thus, emission per unit star

formation (L/ψ)template ratio is lower for M82 and thus for our starburst template. This ratio

provides the proportionality constant in our scaling law L = (L/ψ)templateψ. In this sense

79

therefore, starbursts are less efficient IC emitters per unit star formation. The major sources

of uncertainty for starbursts are the magnetic field strength, and the injection e/p ratio–if it

is indeed different for starbursts as suggested by Lacki et al. (2012).

3.6 Results: Inverse Compton Contribution to the

Extragalactic Background

Having established a one-zone galaxy template, and explored physically plausible variations

in it, we proceed to compute the IC contribution to the EGB measured by Fermi-LAT from

star-forming galaxies. This is given by the well-known expression

IE =c

(1 + z)

dt

dz

Lγ[(1 + z)E, z] dz (3.27)

with |dt/dz| = (1+z)−1 [(1+z)3Ωmatter +ΩΛ]−1/2H−10 . Here we assume a flat ΛCDM universe

with H0 = 71 km s−1 Mpc−1, Ωmatter = 0.3, and ΩΛ = 0.7, following Fields et al. (2010).

The cosmic IC luminosity density (i.e,. cosmic volume emissivity) is given by substituting

our single-galaxy luminosity of eq. (3.25) into the luminosity function of eq. (3.2). Because

we have Lic ∝ ψ, the IC emission of a star-forming galaxy traces its star-formation rate.

Thus, the IC luminosity function dngal/dLγ at each redshift is proportional to the luminosity

function of any tracer of star formation at that redshift. We explicitly distinguish for both

normal and starbursts over cosmic time. Suppressing the energy dependence for clarity, we

have

Lγ(z) =

Lγdngal

dLγ

dLγ

=Lnorm(z)fnorm(z)

ψMW

ψnormdngal

dLγdLγ +

Lsbg(z)(1 − fnorm(z))

ψM82

ψsbgdngal

dLγdLγ

=ρ⋆,norm(z)

ψMW

Lnorm(z) +ρ⋆,sbg(z)

ψM82

Lsbg(z) (3.28)

80

Fiducial Normal Galaxy modelGalactic Center Normal Galaxy modelFiducial Starburst Galaxy modelTotal IC : Normal + Starburst

Figure 3.3: The inverse-Compton EGB intensity from star-forming galaxies as a functionof energy. The solid blue curve represents our fiducial IC model summing contributionsfrom normals and starbursts shown in dashed blue and green respectively using single galaxymodels of each shown in Fig. 3.2 where interstellar densities evolve with redshift. The dottedblue curve is for an evolving single normal galaxy template with an ISRF corresponding tothe Galactic centre model. The red curve shows the pionic signal from normal star-forminggalaxies only as in Fields et al. (2010). The dashed black curve shows the total contributionfrom normal galaxies ignoring starbursts. The dashed green shows IC contribution of M82type starbursts alone. It is a factor of a few less than that from normals. The solid blackcurve shows the sum total of normal pionic and total star forming IC.

81

where the ρ⋆ = 〈ψ ngal〉 =∫

ψ dngal/dLγ dLγ is the cosmic star-formation rate. Here fnorm(z)

is the fraction of norml star-forming galaxies at redshift, and 1 − fnorm(z) the starburst

fraction. We calculate this as in (Fields et al., 2010), where starbursts are taken to be

galaxies with a star-formation surface densities Σ⋆ > 0.4 M⊙ yr−1 kpc−2 (Kennicutt, 1998);

this gives fnorm(0) ≈ 1 at z = 0, declining as redshift increases, with fnorm(1) = 0.68 at z = 1.

Following Fields et al. (2010) we adopt the Horiuchi et al. (2009a) cosmic star-formation rate,

and a Milky Way star-formation rate ψMW = 1.07M⊙/yr (Robitaille & Whitney, 2010). For

M82, we use ψM82 = 8.0M⊙/yr, consistent with Sanders et al. (2003); Lacki et al. (2012).

The overall amplitude of the cosmic IC luminosity density is thus linearly proportional

to the cosmic star-formation rate. Of course, the detailed spectral shape of the luminos-

ity density reflects redshift dependence of the interstellar matter and radiation encoded in

LMW(z).

Figure 3.3 shows the IC contribution to the EGB for some choices of the Milky-Way

spectrum LMW and also the fiducial M82 template, LM82. We focus first on the results for

the fiducial model, which appear as the solid black curve, and incorporate the full redshift-

dependence of interstellar density encoded in LMW(z). A fitting function for the fiducial

model appears in Table 3.3, valid over the Fermi energy range. For comparison, Fig. 3.3

shows the pionic contribution in red. In our fiducial model, the IC contribution is about

10% of the pionic contribution at the ∼ 300 MeV peak in E2I; growing to about 20% at 10

GeV. This is consistent with our expectations from the order-of magnitude calculations of

§3.2.

The shape of the fiducial IC curve is rather smooth, and rather flat in E2I. This is

because the spectrum it is a redshift-smeared version of the single-galaxy IC emission. This

stands in contrast to the pionic curve that displays its characteristic peak and a steep ∼ E−sp

dropoff at large energies, with sp = 2.75 the propagated proton spectral index.

The bulk of the ficucial IC signal comes from redshifts at the peak of the EGB integrand

in eq. (3.27). This occurs at zpeak ∼ 1. A comparison of the corresponding curves in Figs. 3.2

82

and 3.3 shows that the IC peak in the galaxy luminosity rest-frame energy Eem ≃ 100 MeV

is transformed into an EGB IC peak at ≃ 40 MeV. This translation in energy space is

consistent with redshifting by a factor 1 + zpeak.

IC (Fiducial) : Normal Only

Figure 3.4: A summary of the star-forming galaxy contributions to the EGB intensity, shownas a function of energy. The blue points represent the Fermi-LAT data points (Abdo et al.,2010k). The solid red and blue curves show the gamma ray intensity due to pionic andfiducial IC components from star-forming galaxies as in Fig. 3.3. The solid black curverepresents the sum of the two. The shaded yellow band gives our estimate of the uncertaintyin the predicted total signal, which is dominated by systematic errors that are common tothe IC and pionic components.

Our fiducial model sums the emission from normal and starburst galaxies; the contribu-

tions from each of these star-formation modes are shown separately in Fig. 3.4. The dashed

blue curve shows the EGB signal from normal galaxies, which we see dominates the total IC

contribution. The IC component from starbursts is shown in dashed green, and is lower than

the normal galaxy contribution by a factor ∼ 2− 4. This suppression arises for two reasons.

First, we found that the starburst photon output (L/ψ)template per unit star formation is

83

lower than that of normal galaxies. In addition, the fraction of galaxies that are starbursts

is small today, and while this fraction grows with redshift, it is still subdominant at z ∼ 1

where the EGB contribution peaks.

The dashed black curve in figure 3.3 represents the total contribution of normal galaxies

to the EGB, summing their pionic and IC contributions. Adding the starburst IC model

leads to the solid dashed curve showing minor increase at the energies beyond 100 GeV. Now,

this computation does not include contribution of pionic emission from starbursts, which has

large uncertainties, (eg., Lacki et al., 2012; Paglione & Abrahams, 2012). Indeed, some have

argued that pionic emission from starbursts makes a small contribution to the EGB (e.g.,

Stecker & Venters, 2011), some models predict starburst contributions should saturate the

observe EGB (e.g., Lacki et al., 2012). It is worth mention that the uncertainties in the

starburst IC contribution are smaller than in the pionic case, largely because the electron

propagation is controlled by calorimetry in both the normal and starburst cases. Of course,

the starburst IC contribution is still rather uncertain, particularly if the the e/p injection

ratio in starbursts is much higher that observed in the Milky Way (Lacki et al., 2012; Paglione

& Abrahams, 2012).

Figure 3.2 includes curves which illustrate the parameter sensitivy of our IC results. To

give a sense of the effect of the adopted ISRF, we have computed the IC EGB resulting from

an evolving ISRF whose spectral shape is appropriate for the Galactic centre. This model

corresponds to the dotted blue curve in Fig. 3.2(a). Here again, the main result is that the

curves are very similar. In detail, the Galactic centre has a slightly higher ratio Uisrf/Umag of

starlight-to-magnetic energy density. Thus, electron energy losses are more IC dominated,

and finally the overall IC flux is closer to calorimetric and so slightly higher.

Our fiducial model assumes the ISRF components (other than the CMB) are dominated

by short-lived stars and thus have a redshift evolution that scales with the cosmic star-

formation rate. An extreme alternative is that the optical ISRF is dominated by long-lived

main sequence stars whose energy density scales with the stellar mass and thus the integrated

84

cosmic star-formation rate, as in eq. (3.11). Adopting this scenario, we have calculated the

IC contribution to the EGB. For clairty, we have omitted the curve from Fig. 3.3. but it

is somewhat smaller than the fiducial model across all energies. This is easily understood:

at high redshifts the star-formation rate is higher than today, but the stellar mass is lower

and thus offers less optical photons and reduced calorimetry. But the effect is not dramatic;

even in this extreme case, the reduction in the IC signal is less than a factor of two relative

to our fiducial model.

Thus, we see that the IC results do depend on the adopted interstellar densities and

their redshift evolution. But for reasonable choices, the variations in the final IC EGB

result are ≈ 10±0.3. The IC calculation in this sense appears rather robust to the systematic

uncertainties in the modeling.

It is useful to compare the IC luminosity density in eq. (3.28) with that for pionic emission.

As noted above, the cosmic IC emissivity is proportional to the cosmic star-formation rate,

along with a weak sensitivity to the redshift evolution of interstellar densities. In contrast,

the Fields et al. (2010) pionic emission depends on the product of cosmic star-formation rate

and mean interstellar gas mass, via the nonlinear scaling Lπ ∼ ψ1+ω with ω = 0.714. Thus,

the pionic emissivity represents a different moment of the cosmic star-formation rate Lπ =

〈Lπngal〉 ∼ 〈ψ1+ωngal〉. This nonlinear moment gives different results depending on how the

star-formation distribution evolves with redshift. Two extreme limits correspond to: (a) pure

density evolution, wherein galaxies have constant star-formation rates but evolving number

density; versus (b) pure luminosity evolution in which the star-forming galaxy density is

constant but the star-formation rates all evolve with redshift. The pionic curves we show

correspond to the fiducial Fields et al. (2010) pure luminosity evolution case, which gives a

contribution to the EGB larger by a factor ∼ 3 than in the pure density evolution case. In

the IC calculation, the pure luminosity and pure density evolution are degenerate, in that

these models all give the same cosmic star-formation rate and thus the same IC output.

Figure 3.4 represents the central result of this paper, summarizing the EGB predictions

85

for gamma-ray emission from star-forming galaxies in the Fermi energy range. We show the

fiducial IC curve from Fig. (3.3), the fiducial normal-galaxy pionic model of Fields et al.

(2010), and the total EGB intensity that sums these components. Over the 100 MeV to

300 GeV range, the pionic component dominates the normal galaxy contribution. But as

we have seen, the IC curve is a much flatter function of energy. The IC component thus

becomes systematically more important at higher energies. Consequently, high-energy slope

of the total emission is less steep than that of the pionic signal, and closer to the Fermi-LAT

data shown in blue (Abdo et al., 2010k).

Including the IC emission thus improves somewhat the agreement between star-forming

prediction and the observed EGB slope, a weaker point of the Fields et al. (2010) pionic-

only model. Up to about ∼ 10 GeV, the central values of the model fall below the data,

but are consistent within the theoretical and observational error budget (discussed below).

At higher energies, the model underpredicts the data. Our pionic model neglects starburst

galaxies; it may well be that their pionic emission is important at this energy range (e.g.,

Lacki et al., 2011). In addition, other unresolved sources (Strong et al., 1976a,c) such as

blazars (e.g., Padovani et al., 1993; Stecker et al., 1993; Mucke & Pohl, 2000; Stecker &

Venters, 2011) and other active galaxies (e.g., Grindlay, 1978; Kazanas & Protheroe, 1983),

starbursts (Thompson et al., 2007) are also guaranteed to play a role . In any case, star-

forming galaxies clearly are an important contribution to the Fermi EGB signal, and could

well be the dominant component.

Note also that the IC curve does not become important at low energies until far below

the pion bump. Thus, the Fields et al. (2010) conclusions stand: the pion feature should

still remain as a distinguishing characteristic of a significant star-forming contribution to the

EGB. Measurements below ∼ 200 MeV should show a break in the EGB slope if star-forming

galaxies play an important role. Such a spectral feature provides one way to discriminate

between star-forming galaxies and other sources, such as the guaranteed contribution from

unresolved blazars.

86

Table 3.3: Fitting Functions

Form: Y = 10aX3+bX2+cX+d, with X = log10(E/1 GeV)Y a b c d

E2emLnormal(Eem) [GeV s−1] −2.51 × 10−2 −5.03 × 10−2 −9.49 × 10−2 40.246E2Inormal(E) [GeV cm−2 s−1 sr−1] −1.53 × 10−2 −7.38 × 10−2 −0.130 −7.63

E2emLstarburst(Eem) [GeV s−1] −9.16 × 10−3 −7.38 × 10−3 −5.41 × 10−2 40.888E2Istarburst(E) [GeV cm−2 s−1 sr−1] −2.37 × 10−3 −2.59 × 10−2 −6.29 × 10−2 −7.86

Our estimate of the uncertainty in the total normal galaxy emission is represented by the

shaded band in Figure 3.4. The errors in the pionic contribution dominate, and are taken

from Fields et al. (2010). For the IC contribution, the errors are dominated by systematic

uncertainties in the Milky-Way star formation rate ψMW and in the normalization ρ⋆(0) of

the cosmic star-formation rate that are both ∼ 40% (Fields et al., 2010). These factors are

common to the overall normalizations of both the pionic and IC signals, cf. eq. (3.28). The

effect of the uncertainties in the parameters magnetic field and interstellar densities on the

normal IC emissivity is estimated from variation in the curves in Fig. 3.2(a) to be ∼ 50%. We

propagate the errors in IC and pionic emissivities and the common errors in star formation

rates, to construct the yellow band in Fig. 3.4. Because the pionic signal dominates the

total, it also dominates the total error, and the uncertainty in the IC calculation does not

change the result much except at the very highest energies where other systematics may be

important.

We have ignored the effect of intergalactic absorption of the high-energy gamma rays

via γγebl → e+e− photo-pair production in collisions with extragalactic background light

(e.g., Salamon & Stecker, 1998; Abdo et al., 2010e; Stecker et al., 2012). This attenuation

starts to become significant for gamma rays over few tens of GeV. Therefore, beyond these

energies our results will be suppressed by amounts that depend on the optical depth for these

high-energy gamma rays. In this context it is worth noting that Fermi-LAT observations

of the z = 0.9 active galaxy, 4C +55.17 do not show significant absorption of gamma rays

87

up to about 100 GeV (McConville et al., 2011). This object lies within the z ∼ 1 regime

where most of the IC contribution to the EGB occurs; thus we might expect that attenuation

should be mild at Fermi energies. Recall also that at high energies, the electron propagation

should include catatrophic losses from IC interactions, which have not been included in our

calculations.

3.7 Discussion and Conclusions

We have calculated the contribution to the extragalactic gamma-ray background due to

inverse-Compton emission from star-forming galaxies. To do this we model the IC emission

in individual star-forming galaxies, which arises from cosmic-ray electron interactions with

the ISRF. For normal star-forming galaxies, we normalize to the present-day Milky Way IC

luminosity as determined by GALPROP (Strong et al., 2010). For starbursts, we use the M82

template guided by observations (de Cea del Pozo et al., 2009) and other results (Lacki et al.,

2012; Paglione & Abrahams, 2012). We also provide a prescription for redshift evolution of

interstellar matter and energy densities, based on equipartition arguments. Assuming cosmic

rays are accelerated by supernovae implies that a galaxy’s IC luminosity scales as Lic ∝ ψ;

this further implies that the cosmic IC luminosity density or volume emissivity is proportional

to the cosmic star-formation rate: Lγ ∝ ρ⋆.

This linear dependence of IC luminosity with star-formation rate has important im-

plications in light of the star-forming galaxies resolved by Fermi. Namely, the Lic ∝ ψ

trend provides a poor fit for Fermi galaxies, for which the observed correlation is non-linear:

Lic ∝ ψ1.4±0.3 (e.g., Abdo et al., 2010f). This implies that the IC component is subdominant

in Fermi galaxies, adding indirect evidence for the primacy of pionic emission in star-forming

galaxies and prefiguring the trends we find for the EGB.

Turning to the EGB, we find that the IC contribution has a very broad maximum in

E2Iic at E ≃ 40 MeV, falling off very gradually away from this peak. This shape is a

88

redshift-smeared reflection of the underlying Milky-Way-like spectrum from the individual

galaxies. In fact, the IC emission is so broadly peaked that is it effectively featureless. This

contrasts markedly with the distinctive pionic feature from cosmic-ray hadronic interactions

in star-forming galaxies.

The amplitude and shape of the IC contribution to the EGB depends on the nature of

the interstellar radiation and matter fields, and their evolution with redshift. However, we

find that when we adopt different but physically motivated variations to our fiducial model,

the final EGB predictions change relatively little. This rough model-independence of the

IC EGB is a consequence of partial calorimetry, i.e., the fact that IC photons represent

a substantial fraction of cosmic-ray electron energy loss, so that the IC output is tied to

cosmic-ray production via energy conservation.

We find that in all of our models the IC signal is always smaller than the pionic contribu-

tion. However, while this is true, IC also has a substantially flatter spectrum, so that the IC

becomes increasingly important away from the pionic peak. This implies that, at least for

normal star-forming galaxies, the IC contribution should dominate over pionic at high and

low energies. For normal star-forming galaxies, one-zone models for scale heights h < 2 kpc

IC can exceed pionic in a narrow window about TeV energies at the high end and below

100 MeV at the low end. However, at higher energies the opacity of the universe becomes

important and losses become catastrophic, and requires different techniques to handle cor-

rectly. At energies below the Fermi-LAT range, the IC emission will be supplemented by

processes such as bremsstrahlung, which we have not included; a detailed treatment of the

MeV regime appears in Lacki et al. (2012).

The approximate calorimetric relationship between IC photons and cosmic-ray electrons

has important consequences. The main redshift dependence of the cosmic IC emissivity is a

linear dependence on the cosmic star-formation rate. Thus, our results are independent of

whether the cosmic star-formation rate is a result of pure luminosity evolution, pure density

evolution, or something in between. This is in contrast to the pionic case, which depends

89

nonlinearly on the star-formation luminosity function and so breaks the degeneracy between

the pure luminosity and pure density evolution cases.

Our IC results for normal and starburst galaxies share important similarites. In both

cases fractional calorimetry controls the basic scaling with star-formation rate, and leads to

similarly broad spectral features that are not very different from power laws. While the M82

starburst template has a higher gamma-ray luminosity than the corresponding Milky-Way

normal galaxy template, the photon output (L/ψ)template per unit star formation is lower for

starbursts. For this reason, and becuase of the relative rarity of starburst galaxies, the IC

contribution to the EGB is dominated by that of normal galaxies. However, the starburst

model has fewer observational constraints and thus larger uncertainties. The starburst IC

contribution could become very important if, for example, the e/p injectio n ratio is very

large in starbursts Lacki et al. (2012).

To simplify the discussion, the IC and pionic models presented here neglected the effects of

Type Ia supernovae, implicitly assuming instead that all supernovae in star-forming galaxies

are due to core-collapse. Lien & Fields (2012) considered in detail the effect of Type Ia

supernovae on the pionic signal. They found that a self-consistent treatment includes both

the addition of Type Ia explosions as cosmic-ray accelerators, but also as part of the Milky-

Way normalization of the cosmic-ray/supernova ratio. The effects largely cancel, so that in

the end, the net pionic galactic luminosity and EGB does not change substantially. In the

case of IC emission a similar cancellation will occur. The only new contribution of possible

importance comes from Type Ia explosions arising from long-lived progenitors in quiescent

(i.e., not actively star-forming) galaxies such as ellipticals. Lien & Fields (2012) show that

the pionic emission from these systems could be large if they have a substantial reservoir

of hot, X-ray-emitting gas. But supernova rate in these galaxies represents a subdominant

fraction of cosmic Type Ia activity, which itself is substantially smaller than the cosmic core-

collapse rate. Thus, the IC contribution from these systems will be small. And so inclusion

of Type Ia supernovae in a self-consistent way would change our results very little, less than

90

the uncertainties in the model.

There remains room to improve on our model. Future work would benefit from obser-

vational progress in clearly identifying an IC signal from individual star-forming galaxies,

at energies away from the pionic peak. Theoretical work would benefit from a more de-

tailed model for the ISRF and its evolution, and from the use of additional multiwavelength

constraints on the cosmic ray electrons.

More broadly, a solid identification and quantification of the main components of the EGB

remains a top priority for gamma-ray astrophysics and particle cosmology. Extending the

Fermi EGB energy spectrum to both higher and lower energies will provide important new

constraints. At sufficiently high energies, the cosmic opacity due to photo-pair production

must become apparent if the EGB is dominated by any sources at cosmological distances

(e.g., Salamon & Stecker, 1998; Abdo et al., 2010e; Stecker et al., 2012). And at energies just

below those reported in Abdo et al. (2010k), a break in the EGB spectrum is an unavoidable

prediction if the signal is dominantly unresolved pionic emission from star-forming galaxies

(both normal and starbursts). An independent probe of EGB origin lies in anisotropy studies

(Ando & Komatsu, 2006; Ando & Pavlidou, 2009; Hensley et al., 2010; Ackermann et al.,

2012a; Malyshev & Hogg, 2011). Regardless of the outcome, an assay of the EGB components

will provide important new information (and perhaps some surprises!) about the high-energy

cosmos.

3.8 Acknowledgements

It is a pleasure to acknowledge stimulating conversations with Keith Bechtol, Vasiliki Pavli-

dou, Troy Porter, Tijana Prodanovic, Amy Lien, Andy Strong, Todd Thompson, and Toni

Venters. After our paper was first submitted, we benefitted greatly from discussions with

Brian Lacki regarding his similar work. BDF would also like to thank the participants of

the 2012 Sant Cugat Forum on Astrophysics for many lively and enlightening discussions.

91

This work was partially supported by NASA via the Astrophysics Theory Program through

award NNX10AC86G.

92

Chapter 4

Statistics of gamma-ray sources

This chapter is an outlines a work in progress with co-authors Eric Baxter (first author),

Scott Dodelson and Brian Fields.

There is great debate about the relative contribution of various gamma-ray emitting

sources to the diffuse EGB. The way to resolve this is to study the various distinguishing

properties of these sources as their gamma-ray spectral slope ,angular anisotropy in the

source distribution, correlation with other wavelengths, etc. The statistics of the photons

themselves bear a signature of the sources producing them. There are a number of candidates

that contribute to the diffuse EGB. AGNs, star forming galaxies, pulsars, etc. are examples

of known astronomical sources. In addition, there is possibility of unknown or exotic sources

contributing to the EGB. The most exciting possibility is the contribution from particle

dark matter via decays and annihilations producing gamma-rays. The known astronomical

sources would be foregrounds to this. Thus it is, crucial to model them accurately not only

in order to understand them but also to seperate them from a possible signature of dark

matter.

For e.g., AGNs are fewer and brighter individually. Amongst these are blazars, AGNs

whose jets point towards us. They are the dominant source class in the gamma-ray emitting

point source catalog (Nolan et al., 2012a). Star forming galaxies on the other hand are far

more numerous, but individually dim. Thus, one expects them to have a different photon

distribution. The gamma-ray emissivity, qγ of a distribution of sources can be given as,

qγ = nγLγ (4.1)

93

Blazars have a number density of nγ ∼ 10−9Mpc−3 and an average luminosity, Lγ ∼

103photons sec−1. Compared to this the star forming galaxies on the other hand have

nγ ∼ 10−3Mpc−3 and Lγ ∼ 106photons sec−1. This rough estimate suggests that the emis-

sivity from a distribution of few, bright blazars is comparable to that from a distribution of

numerous, but dim star forming galaxies.

A detailed analysis of statistics of source photon counts distribution however, has the

potential to reveal the true relative contribution of these sources. This is undertaken in

Baxter, Chakraborty, Dodelson and Fields,.(in prep). An method based on extension of the

P(D) analysis described in (Scheuer, 1957) is used to analyse the photon counts from Fermi

-LAT. The quantity of interest here is the probability Pi(C) of finding Ci photons in the

pixel i of the sky. The individual Pi(C) of the different contributing source types here are

determined from their redshift-dependent luminosity functions. The luminosity functions of

five different source types including the two classes of gamma-ray emitting blazars, are used

namely

1. Star forming galaxies (SFGs)

2. Flat Spectrum Radio Quasars (FSRQs)

3. BL Lacertae Objects (BL LACs)

4. Dark matter annihilation from Galactic subhalos

5. Diffuse emission from the Galaxy

4.1 Formalism

From the luminosity function dndL

of a source type, the differential source flux distribution

dndF

is computed. This is an observed quantity. Of course, what is really computed is the

94

probability of observing flux F, from a single source along the line of sight, P1(F ) as,

P1(F ) ∝∫

dzdLd2

L

(1 + z)2E(z)

dnc

dLδ

(

F − L(1 + z)−(α−2)

4πd2L

)

(4.2)

=

∫ z(Lmax,F )

z(Lmin,F )

dzd4

L(1 + z)(α−2)

(1 + z)2E(z)

dnc

dL

L=4πd2L

F (1+z)(α−2)

(4.3)

where dL is the standard luminosity distance. And the delta function ensures that the flux is

F for a source of luminosity L at dL. The underlying assumption here is that sources have a

power law spectral energy distribution, an assumption good enough for star forming galaxies

and blazars in the regime of interest. z(Lmin, F ) and z(Lmax, F ) are the redshifts expressed

as functions of the minimum and maximum, luminosity Lmin and Lmax, respectively, that

produce observed fluxes F . The source flux distribution dndF

is related to this by,

dn

dF= µP1(F ) (4.4)

where µ is the mean number of sources per unit solid angle along the line of sight.

Having determined P1,a for each source type a, the probability of measuring a sum total

flux F of all the source types, i.e. the flux probability distribution is the convolution of the

individual P1,a.

P (F ) = P1,a ⋆ P1,b..... ⋆ P1,z (4.5)

In reality LAT observations constitute discrete photon counts. The continuous flux probabil-

ity is related to the discrete photon count probability in a particular pixel i via the exposure

in that pixel as

Pi(C) =

dFPi(F )exp(−EiF )(EiF )C

C!(4.6)

and analogous to the continuous flux distribution of a number of sources as in eq. (4.5), the

95

discrete counts distribution is also a convolution of the individual source types,

Pi(C) =C∑

J,K,L,M=0

P FSRQi (J)P SFG

i (K)PDMi (L)PBLLac

i (M)

PGali (C − J −K − L−M) (4.7)

Thus, the data i.e. the counts distribution puts constraints on the model parameters in the

luminosity function of the source types.

96

Chapter 5

High Energy Polarisation

The work in this chapter was done in collaboration with co-authors Vasiliki Pavlidou and

Brian Fields. It is yet to be submitted to a journal.

5.1 Introduction

Active galaxies or AGNs (active galactic nuclei) are some of the most luminous yet mysterious

astronomical objects in the universe. Their particle and radiative emissions are powered by

central supermassive black hole accreting matter. Part of the gravitational energy associated

with accretion is converted into energy of particles such as cosmic rays and neutrinos and

high energy radiation like X-rays and gamma-rays. Thus, these particles and radiation are

messengers of the extreme astrophysical conditions in the core of active galaxies.

Blazars are distant AGNs where the observer’s line of sight is along the jet axis, i.e.

the observer looks down into the jets. Polarised, high energy photons are expected to be

produced by inverse-Compton scattering of low energy seed photons in blazars (McNamara

et al., 2009; Krawczynski, 2012). Seed photons can come from external sources (Meyer et al.,

2012) like the accretion disk and cosmic microwave background (CMB) and internal sources

within the jet like the synchrotron photons (Zacharias & Schlickeiser, 2012) emitted by the

electrons. External target photons through external Compton (EC) mechanism are expected

to produce a lower degree of polarisation than internal photons such as self-synchrotron

Compton (SSC) (McNamara et al., 2009; Krawczynski, 2012).

Various properties of the radiation from blazars like the overall intensity, spectrum, and

97

variability have been studied with multiwavelength observations. However, polarisation at

high energies has received much less attention, not just for blazars, but for any source

class. High energy polarisation measurements of the Crab nebula / pulsar (e.g., Dean et al.,

2008) are the most significant of the high energy polarimetric observations with the Crab

being used for calibrating polarisation observations. There’s also evidence for solar flares

producing polarised X-rays (McConnell et al., 2003). Amongst transient sources, there are

very few successful observations of gamma-ray bursts (GRBs) as listed in Chang et al.

(2013) that . No high energy polarimetric data exists for blazars. This is in part due

to the challenges in measurement of polarisation in X-rays and soft gamma-rays. However,

with numerous X-ray and soft gamma-ray polarimeters at various stages of planning, design

and operation and studies of optical / FIR polarisation properties of blazars underway, a

systematic study of high energy polarisation from blazars is due. Thus, the amount and

degree of high energy polarisation from blazars is an empirical and open question. However,

the theoretical expectations are described in section 5.3. Throughout when referring to high

energy polarisation, we mean X-ray and soft gamma ray polarisation.

Of course, polarised emission from GRBs and blazars at low energies has been studied in

a lot more detail (e.g., Lazzati, 2006; Rossi et al., 2004; Wardle & Kronberg, 1974; Agudo

et al., 2010; Fujiwara et al., 2012). Synchrotron emission is intrinsically polarised (Rybicki

& Lightman, 1986) extending over a broad range of energies and is the main emission mech-

anism in GRB prompt and afterglow emission as well jet emission from both GRBs and

active galaxies. The polarisation depends on magnetic field structure in the environment as

well as the energy distribution in the jets (Lazzati, 2006). Blazars have two broad peaks in

the spectral energy distribution of the jet emission. The first, lower energy peak is due to

synchrotron emission that is intrinsically polarised (Rybicki & Lightman, 1986). It ranges

from radio to as high an energy as X-rays. Observed emission from blazars at lower energies

show polarisation (e.g. Fujiwara et al., 2012). The second peak is typically associated to

inverse-Compton scattering of low energy photons that retains a fraction of this polarisa-

98

tion due to the nature of the inverse Compton scattering process (Bonometto et al., 1970;

Krawczynski, 2012).

Polarisation is a key ingredient in the multimessenger understanding of blazars. The

gains of studying this property are enormous. Polarisation can distinguish between different

emission mechanisms of blazars such as different leptonic models and potentially between

hadronic and leptonic. The connection with the hadronic models is mainly through the

secondary leptons. Hadronic models are accompanied by neutrino emission, whereas polari-

sation should predominantly be from leptonic models via Compton scattering. So there may

be an interesting neutrino polarisation connection to explore. Existence of polarisation in X-

ray and gamma-ray emission from blazars distinguishes it from unpolarised sources at these

energies. Furthermore, if blazars were to possess a reasonably high degree of polarisation,

being high redshift sources like gamma-ray bursts, they could be used for probing Lorentz

invariance violation using vacuum birefringence i.e. the rotation of the plane of polarised

light due to quantum gravity models (e.g., Coleman & Glashow, 1999; Ellis et al., 2000;

Laurent et al., 2011). The variability of blazars is a plus which can serve as a diagnostic for

monitoring the relation between the jet activity and polarisation. This will also allow mea-

surement of polarisation during the flaring states of those blazars that would under normal

circumstances be undetectable in polarisation.

In this paper, we wish to focus on detection prospects of X-ray and soft gamma-ray

polarisation of blazars with polarised seed photons in their jet. In Sec. 5.2, we discuss the

degree of polarisation expected from inverse-Compton scattering of polarised low energy

photons by relativistic electrons in the jet with a power law distribution. The minimum

detectable polarisation (MDP) for various telescopes in general is reviewed with detectability

of a typical blazar in Sec. 5.3. lists some of the brightest blazars and the time required to

detect the MDP for different polarimeters. The detection prospects are improved due to

flaring as described in section 5.4. In the section 5.5, we discuss the potential of polarised

data to indirectly constrain hadronic emission. And in section 5.6, the conclusions along

99

with future directions are described.

5.2 Degree of polarisation of blazar

The degree of polarisation is the fraction of polarised light. Polarisation may be linear,

circular or more generally elliptical. In general, the extent of polarisation is quantified in

terms of the Stokes’ parameters, I, Q, U, V where as usual

I = Ix + Iy

Q = Ix − Iy

U = Ia − Ib

V = Ir − Il (5.1)

a, b represent directions at 45 to x, y respectively. I is the intensity of light, Q and U are

measures of linear polarisation along any direction say x and along 45 from x respectively

(along a) and V is the measure of circular polarisation. Q = 100% implies that the light is

polarised along the + or - x-axis. If Q = U = 0, then V = I and light is circularly polarised.

The degree of polarisation is in general given by,

Π =

Q2 + U2 + V 2

I(5.2)

The degree of linear polarisation consists only of Q and U, Πlin =

√Q2+U2

I. Bonometto et al.

(1970) calculated the polarisation of photons produced by inverse-Compton scattering of

an arbitrary distribution of photons off an arbitrary, distribution of electrons. The simpler

case of monochromatic beam of photons is studied that sets up future papers to study more

realistic and complicated distribution of photons. The electrons are unpolarised. Both the

target photons and the electrons are assumed to not have any spatio-temporal variations.

There are two key conditions central to this calculation relating the initial ǫ and scattered

100

photon energies Eγ. The first is the definition of inverse-Compton scattering,

Eγ ≫ ǫ (5.3)

The second is the Thomson limit in the center of mass frame, valid for a lot number of

astrophysical scenarios including that for blazars,

me

ǫ

me

≪ 1 (5.4)

For the inverse-Compton upscattered radiation, Bonometto et al. (1970) find the degree of

linear polarisation for completely (100%) polarised target photons is computed to be

ΠBCS =Σ1 + Σ2

Σ1 + 3Σ2(5.5)

where the Σ’s are related to the electron distribution function, m(γ) as a function of the

Lorentz factor, γ,

m(γ) =dne

dγγ−2 ; Σ1 =

∫ 1

0

m(γ)

(

x2 − 1

x2+ 2

)

dx ; Σ2 =

∫ 1

0

m(γ)(1 − x)2

x2dx (5.6)

ne is the number density of electrons and x = γmin/γ, with γmin being the minimum Lorentz

factor required by the electron to produce an up-scattered photon of a given energy given

by γmin = Eγ

2 me(1 +

1 + 4 m2e

ǫEγ). Furthermore, their results are computed for an isotropic

distribution of electrons. This is a good assumption as the relativistic electrons under consid-

eration here scatter photons into a narrow cone ∼ γ−1 around the direction of propagation.

And as long as the electron distribution doesn’t vary a lot over this solid angle in the rest

frame of the electron plasma, the results hold true.

According to Bonometto et al. (1970), unpolarised target photons lead to unpolarised

light post inverse-Compton scattering. This can be understood in terms of their observation

101

that Stokes’ parameters change signs in changing from one polarisation state to its orthogonal

state as evident from eqns 7.3 − 7.5 in their paper. As a result, unpolarised target photons

that can be viewed as a superposition of equal proportions of orthogonal polarisation, cancel

out to give zero net polarisation. This is in contrast to Thomson scattering, that can produce

polarisation from unpolarised photons.

Krawczynski (2012) re-evaluates analytically and numerically the calculation of polari-

sation due to inverse-Compton scattering. The numerical calculation is required to unam-

biguously ascertain the range of validity of results and also for precision in determining the

degree of polarisation. The degree of polarisation is explicitly computed for the case of

photons scattered in the jet of blazars verifying the result that unpolarised target photons

lead to unpolarised scattered photons. The self-synchrotron and external Compton cases are

studied explicitly with the conclusion that the external Compton leads to a smaller degree

of polarisation owing to low degree of polarisation of target photons either from the start as

for the CMB or by averaging over axisymmetric configuration in case of broad line region

clouds. The degree of linear polarisation from scattering of the mono-energetic unidirectional

photons off a power-law distribution with index p of electrons is given by,

Πpl =(1 + p)(3 + p)

(11 + 4p+ p2)(5.7)

For p = 2, Πpl ≈ 65% and p = 3, Πpl ≈ 75%. Now, Compton scattering of polarised

light leads to polarised light, and the degree of polarisation Π, is modified from the initial

polarisation, Πin, by this above mentioned factor as (Bonometto et al., 1970; Krawczynski,

2012) for a non-thermal electron distribution,

Π = Πpl × Πin (5.8)

Theoretically, for a power-law distribution of electrons, the degree of polarisation of syn-

102

chrotron emission is given by (Rybicki & Lightman, 1986),

Πsync =p+ 1

p+ 7/3(5.9)

that would give a large values for typical p = 2 − 3. However, various effects like Faraday

rotation and inhomogeneities in the magnetic field can reduce this value 1. In light of

these complications, it is more reliable to use the observed values of Πin. We thus, plan

to test empirically the relation between degrees of low energy polarisation and high energy

polarisation. From Fujiwara et al. (2012), more than half of the gamma-ray blazars are

highly (> 10%) polarised for at least part of the duty cycle. This quick estimate suggests,

that from Krawczynski (2012) and Fujiwara et al. (2012), a degree of polarisation of a few %

at least, for Fermi -LAT blazars in X-rays and soft gamma-rays. According to simulations

in McNamara et al. (2009), the numbers are even higher. However, we will use the results

from Krawczynski (2012) based on the analytical calculations in Bonometto et al. (1970).

These numbers are also more conservative.

Now in general, Π is a function of the direction of target photons and energy of the

scattered photons. Figures 8 and 9 in Krawczynski (2012) show Π to be an increasing

function of the scattered photon energy implying by extrapolation that soft gamma-rays

are likely to be polarised. However, a detailed calculation appropriate at hard X-ray / soft

gamma-ray energies is needed to show this explicitly. Also, the calculation of Krawczynski

(2012) is for uniform distributions of magnetic field, seed photons and electrons. However, in

realistic jet models, a full calculation including their spatio-temporal variation as suggested

by eq.(55) in Krawczynski (2012) should be performed.

1http : //www.astro.uni − wuerzburg.de/ mkadler/scripts/jets ss12/jets hand04.pdf

103

5.3 Minimum Detectable Polarisation (MDP) -

Detection sensitivity of telescopes

In terms of classical electrodynamics, polarisation is defined in terms of the direction of the

electric field vector in an electromagnetic wave. A polarisation analyser is a component of

a polarimeter, that passes light with polarisation vector aligned to a particular direction

or axis, and blocks the light with an orthogonal polarisation detector. Of course, for high

energy radiation as for soft gamma-rays and X-rays, the quantum mechanical picture of

individual photons is more applicable. In the quantum picture, it is the helicity of photons

i.e. the projection of the photon’s spin along the direction of propagation which defines

polarisation. With spin 1, photons have three polarisation states, +1, 0, -1. Of these, only

the transverse components corresponding to the ±1 states are non-zero for a real, massless

photon. These correspond to the right and left handed helicities or equivalently the right

or left circular polarised light. Of course, independent of the basis in which polarisation

is expressed, it is measured by detectors in terms of differences of intensities quantified by

Stokes’ parameters given above in eq. (5.1). Measurements of intensities at different angles is

the basic technique behind measuring polarisation. There are several processes which record

and retain the angular variation of intensity of polarised light such as Bragg reflection,

Compton scattering, photoelectric effect and pair production (Novick, 1975) 2. Of these,

the Compton polarimeters are appropriate for the energy range of photons discussed here,

wherein, the full Klein-Nishina cross-section depends on the angle between the polarisation

vectors of initial and scattered photons (Heitler, 1954).

Whether in the classical picture or the quantum picture, the angular dependence of the

polarisation signal can only be quantified in a statistical sense due to presence of backgrounds

and noise. According to Novick (1975); Weisskopf et al. (2010), for a source producing N

counts, the probability distribution of fractional polarisation say P due to statistical fluctu-

2http : //www.cesr.fr/spip.php?article863

104

ations, measured by an analyser passing one polarisation state, f(Π) = exp

(

−NΠ2/4

)

.

Now a real detector has backgrounds and thus, the polarisation signal is a modulation

a = δΠ over and above the background. Therefore, the probability of measuring this po-

larisation signal a, due to the source in presence of the background is given approximately

by, f(Π) = exp

(

−N(Π − a)2/4

)

. Of course, the polarisation signal will likely have a

modulation phase, which complicates the probability distribution further (Weisskopf et al.,

2010).

Thus, (Weisskopf et al., 2010), polarisation can be manifested as a modulation of a signal

observed in the azimuth φ about the line of sight in direction of the source. The modulation

occurs at an angular frequency of two times the rotation frequency. The signal-to-noise ratio

for N counts in case of Poisson noise is 1/√N . Therefore, amplitude of modulation over a

signal of RS counts sec−1 in presence of a background RS counts sec−1, that has only 1%

chance of being exceeded, or at 99% confidence is (Novick, 1975; Weisskopf et al., 2010),

a1% =RS +RB

RS

nσ√N

(5.10)

where nσ = 4.29 for 99% confidence level. The detector response to polarised light is quan-

tified in terms of modulation factor, µ which is the amplitude for 100% polarised light in

absence of background. Therefore the detectability of a given source of certain strength by

a given detector is quantified in terms of the minimum degree of polarisation it can detect

called the minimum detectable polarisation. The MDP given a source in general is therefore

given by Novick (1975); Weisskopf et al. (2010); Kalemci et al. (2004),

MDP =nσ

µ RS

(

2(RS +RB)

T

)

(5.11)

with T being the exposure time.

Now given a detector with a certain modulation factor µ and background RB counts sec−1,

105

eq. (5.11) can be inverted to determine the amount of time required to observe a particular

source producing RS counts sec−1. As described above, blazars are theoretically expected

to have polarised synchrotron emission. However, the exact theoretical computation is com-

plicated by effects like Faraday rotation and inhomogeneities in the magnetic field. These

computations are important and need to be performed, but will not be the focus of this

work. Instead we chose to rely on observations.

The GEMS White Paper Krawczynski et al. (2013) proposes for GEMS like missions a

strategy similar to what we describe below. They emphasise on using archival data on low

energy polarisation in order to motivate observations of high energy polarisation. Their idea,

of course is restricted to GEMS like missions. The study here extends to other polarimeters.

And in addition to archival or published data, we plan to perform synergistic polarimetric

observations at low and high energies of blazars like the ones mentioned below using RoboPol.

3

Blazars are observed to show polarised emission at low energies like at IR wavelengths

(Fujiwara et al., 2012). These blazars are selected from the Fermi -LAT catalog, that is an

excellent all sky representation of blazars. We use these observed polarisation values and

determine the degree of high energy polarisation of a blazar as described in section 5.2. We

invert the above eq. (5.11) to determine how long it would take to detect the chosen blazar.

This exercise is performed for 3 of the brightest blazars with the highest degree of near-IR

polarisation. Indeed, there are several other candidates, however, 3C 454.3, PKS 0048-09

and PKS 1124-186 are the brightest blazars that have the highest measured IR polarisation

from the list in and recorded for various polarimeters in Fujiwara et al. (2012). In addition

a source like 3C 454.3 has shown significant variability Wehrle et al. (2012); Jorstad et al.

(2010). Its flaring behaviour makes it a prime candidate for polarisation measurements as

we will comment on in the next section 5.4.

The required observing times for the various X-ray and soft gamma-ray polarimeters

3http : //robopol.org/

106

are computed for these chosen blazars. The polarimeters are PoGOLITE (25 − 100 keV),

GEMS-like mission (2 − 10 keV), ASTRO-H (50 − 200 keV), POLARIX (2 − 10 keV) and

GAP (50 − 300 keV). The flux of the blazars in units of CGS and mCrab (i.e. in terms of

the flux of Crab in that particular range of energies) is tabulated along with the polarisation

of target photons (Πin) and the required observed times in Tables 5.1, 5.2, 5.3, 5.4 and

5.5.

Source Flux Flux Input Polarisation % Observation time(×10−6ergs cm−2 sec−1) mCrab Πin (Flight Times)

3C 454.3 1.85 × 10−4 13.0 10 237PKS 0048-09 4.32 × 10−5 3.04 11.63 3212PKS 1124-186 4.32 × 10−5 3.04 9.45 4861

Table 5.1: Polarisation observational prospects - PoGOLITE: 200 mCrab,10% in 1 flight(20days)

Source Flux Flux Input Polarisation % Observation time(×10−6ergs cm−2 sec−1) mCrab Πin (ksec)

3C 454.3 5.0 × 10−5 3.51 10 3.24PKS 0048-09 5.0 × 10−6 0.35 11.63 239.80PKS 1124-186 9.10 × 10−6 0.64 9.45 2.74 × 103

Table 5.2: Polarisation observational prospects with XRT / ROSAT fluxes - GEMS (2 - 10keV): 1 mCrab, 2% in 1000 ksec

Source Flux Flux Input Polarisation % Observation time(×10−6ergs cm−2 sec−1) mCrab Πin (ksec)

3C 454.3 1.32 × 10−4 9.28 10 214.90PKS 0048-09 4.32 × 10−5 3.04 11.63 1.49 × 103

PKS 1124-186 4.32 × 10−5 3.04 9.45 2.25 × 103

Table 5.3: Polarisation observational prospects with Swift-BAT fluxes - ASTRO-H (50-200keV): 10 mCrab, 10% in 1000 ksec

Source Flux Flux Input Polarisation % Observation time(×10−6ergs cm−2 sec−1) mCrab Πin (ksec)

3C 454.3 5.0 × 10−5 3.51 10 8.03PKS 0048-09 5.0 × 10−6 0.35 11.63 599.42PKS 1124-186 9.10 × 10−6 0.64 9.45 274.90

Table 5.4: Polarisation observational prospects with XRT / ROSAT fluxes - POLARIX(2.0-10.0 keV): 1 mCrab, 10% in 100 ksec

107

Source Flux Flux Input Polarisation % Observation time(×10−6ergs cm−2 sec−1) mCrab Πin (days)

3C 454.3 1.85 × 10−4 13.0 10 4.73PKS 0048-09 4.32 × 10−5 3.04 11.63 97.21PKS 1124-186 4.32 × 10−5 3.04 9.45 64.24

Table 5.5: Polarisation observational prospects - GAP (50.0-300.0 keV): 1 Crab, 20% in 2days

The PoGOLITE results based on our estimates using the analysis in Krawczynski (2012)

suggest that detection of blazars will be challenging with sensitivity of a balloon experiment.

Flaring blazars observed with dedicated balloon flights may have a chance of detecting the

polarisation from blazars. This will demand, planned observations based on triggers from

the optical, UV, IR telescopes monitoring flaring activity. With the future space based

instruments, the sensitivity is much higher and thus, statistically significant detections are

highly likely with reasonable observing times. Now, the space based instruments are not

dedicated merely to polarisation measurements and therefore, observational strategies are

required. Due to the high sensitivity, quiescent or non-flaring blazars are also expected to be

discovered. This is achievable during all sky scans as well as by targeted, longer observations

of the more dim blazars. And for flaring blazars optical, UV, IR triggers can be used for

making detections. These strategies must be kept in mind while planning for future missions.

In addition to using published and archival data, we plan to perform synergistic, mul-

tiwavelength polarimetry with RoboPol 4 and high energy polarimeters. RoboPol is a fast

optical polarimeter designed specifically to study blazars. It is a pointed instrument with

high cadence. The camera onboard the instrument is designed specifically for monitoring of

blazar jets . The polarimeter is mounted on the University of Cretes Skinakas Observatory

1.3 m telescope. From the above results, idea of using a dedicated, specialised polarimeter at

optical wavelengths in conjunction with a high energy polarimeter will significantly improve

the chances of detecting blazars. This is because monitoring the flaring activity of blazars

as described in detail the next section 5.4 can be very useful to improve detection prospects.

4http : //robopol.org/

108

Furthermore, there is evidence that polarisation certainly at low energies is dynamic (Sorcia

et al., 2013). This suggests that given low energy polarisation is ultimately a source of high

energy polarisation, polarimetric variability could be crucial at X-rays and gamma rays too.

Polarimetric variability means that there are times when the degree of polarisation increases

and therefore detectability improves. Also, variability in flux could be connected to that in

polarisation, a connection that would ultimately lead to clues about the underlying mecha-

nisms. Thus, multiwavelength polarimetry sharpen the emission models which connect the

low energy to the high energy emission. Therefore, in order to faithfully probe the emission

models, it is important to do synchronised multiwavelength observations. This is where ded-

icated, specialised like RoboPol can be critical to our understanding of polarisation and in

general emission processes in blazars.

5.4 Flaring

Blazars are known to have flux variations (Nalewajko, 2013). They fluxes can vary by up

to a factor of 10 or 20 when they flare. As discussed in the previous section, flaring activity

of blazars affect the detectability in the polarisation domain. The direct reason is simply

that those blazars that in their quiescent state fall below the detection limit may be pushed

above threshold when they flare. Secondly, the polarisation signal itself may vary due to the

flaring activity. This could in principle reduce the detectability and needs to be determined

observationally. Here we propose a formalism to quantify the former effect, that is the

increase in the detected fraction of blazars due to an increase in their flux or effectively a

lowering of the detection limit by the same factor.

For this we use the formalism of Feldman and Pavlidou (in prep). Every blazar flares

given a long enough time. Feldman and Pavlidou treat the flaring and quiescent blazars as

a single population with two states. Both flaring and quiescent blazars are represented by a

109

single luminosity function. The two states differ by a flaring factor is given by

R =Ff

Fq(5.12)

where Ff and Fq are fluxes in the flaring and quiescent state respectively. This flaring factor

is assumed to be the same for all blazars and uniform in flux.

They find that the source flux distribution fitting the Fermi -LAT data has contributions

from both quiescent and flaring sources. This implies an extrapolation of the distribution

down to the fluxes at keV and MeV energies at which the polarimeters operate. In practice,

this is easy as blazars have a power-law spectrum. And this will be done at the end when

we provide numbers for the fractional increase in the detected blazars.

For simplicity, we demonstrate the formalism in terms of the LAT fluxes above 100 MeV.

Above F ∼ 2× 10−9 photons cm−2 sec−1, the fit to LAT data is in excellent agreement with

the model that has all the blazars flaring. In accordance with Feldman and Pavlidou, if the

duty cycle i.e. the fraction of time a blazar is in a flaring state is χ, then at any given time

the number of blazars per unit flux in the flaring state is related to that in the quiescent

state,

dN

dFq=

1 − χ

χ

dN

dFf= Cq

dN

dFf(5.13)

where Cq = 1−χχ

represents the factor by which the number of blazars in the quiescent state

exceed the number in the flaring state at any given time in the universe. The mean flux 〈F 〉

of blazars, is a weighted average of the flux of blazars in the quiescent state, Fq and that in

the flaring state, Ff .

〈F 〉 = χFf + (1 − χ)Fq (5.14)

Now the all sky source flux distribution is given in terms of the mean flux observed over

110

a long period of time as,

dN

dF= A

(

F

Fb

)−β1

, F ≥ Fb

= A

(

F

Fb

)−β2

, F < Fb

(5.15)

in accordance with (Abdo et al., 2010l). with the parameter values corresponding to the

best fit values for all blazars both FSRQs and BL LACs put together are AFermi = 16.46 ×

10−14cm2 sec deg−2, Fb = 6.60 × 10−8ph cm−2 sec−1 , β1 = 2.49 and β2 = 1.58. The source

flux distribution is normalised here to the break flux, to give A = AFermiF−β1

b = 1.25 ×

105cm2 sec deg−2 ph according to Feldman and Pavlidou.

As a result of flaring, the break flux is shifted to lower value Fb/R. So, the source flux

distribution is modified to include both flaring and quiescent blazars as they may be treated

as a single population given by,

dN

dF

obs

= A

(

F

Fb

)−β1

+ ARCq

(

RF

Fb

)−β1

, F ≥ Fb

= A

(

F

Fb

)−β2

+ ARCq

(

RF

Fb

)−β1

,Fb

R≤ F ≤ Fb

= A

(

F

Fb

)−β2

+ ARCq

(

RF

Fb

)−β2

, F ≤ Fb

R(5.16)

The observed quantity over an integration time T, is of course, the mean flux 〈F 〉. Now,

this leads to two limits, when the observation time is much longer than the period of the

flare, P and vice versa. In the limit that, the observation time is long enough, T ≫ P , all

“flaring events” including multiple flares from the same source are detected. Therefore, the

number of flaring events is given by NuT/P = NuχT/t. In this case, the mean flux is related

111

to the flaring flux as,

〈F 〉 = χFf + (1 − χ)Ff/R (5.17)

=> Ff =R

(R− 1)χ+ 1〈F 〉 = Reff〈F 〉 (5.18)

Reff =R

(R− 1)χ+ 1,

χT

t≫ 1 (5.19)

For typical values of flaring factor, R = 10 and a duty cycle of χ = 1%, Reff = 9.2. For a

longer duty cycle, say χ = 20%, the mean flux remains closer to the flaring flux and therefore

we get a lower Reff = 3.6. Suppose, Nu is the number of sources undetected in the quiescent

state, but that are detected during their flaring state. Therefore,

Nu =

∫ Fsens

FsensReff

dN

dFdF (5.20)

where Fsens is the sensitivity of the polarimeter. Now in principle, computing Nu tells me

the additional number of blazars detected in polarisation. In practice however, from the

fluxes of our candidate blazars in the previous section and the observation times needed to

detect them, it is only the very bright few blazars that will be detected anyway. And so

the bright end of the source flux distribution dNdF

is of interest. Of course, the exact number

of additional blazars detected by the polarimeters will depend on their flux sensitivity (or

MDP) and of course the duty cycle and periodicity of flaring. A comparison of the mean flux

to the sensitivity gives an idea of the increase in detectability. The mean flux is computed

for limit when all blazars originally below the sensitivity, that exceed the sensitivity limit

due to flaring are detected.

Thus, from equation (5.20), the integration range reduces from ∼ 90% to ∼ 72%. Picking

a sensitivity higher than the break flux Fsens = FbReff as most certainly the blazars detectable

are above the break, could potentially almost double of the number of blazars for a duty

112

cycle of ∼ 20%. However, as stated before the actual numbers would depend on a number

of factors including the assumption that all blazars flare. And indeed since there would

be a handful of blazars, instead of a detailed statistical description, individual blazars that

are known to flare and have a high degree of polarisation at the lower energies should be

selected for pointed observations. We expect that the number of blazars detected due to

flaring activity would nearly double.

5.5 Indirect constraints on hadronic emission

As we discussed above, polarisation is a probe of emission mechanisms of blazars. In fact,

polarised X-rays and gamma-rays are primarily a probe of leptonic emission. Amongst

leptonic emission, inverse-Compton scattering by electrons and synchrotron radiation are the

most likely mechanisms of producing polarisation. The blazar spectral energy distribution

consists of two broad spectral peaks or bumps (Abdo et al., 2010m). The lower energy

peak is usually put down to synchrotron emission that is intrinsically polarised (Rybicki

& Lightman, 1986). The higher energy peak is believed to be due to inverse-Compton

scattering of either the synchrotron radiation itself or some other source of photons. As

described above in detail, inverse-Compton emission in blazars is likely to be polarised since,

the underlying target photon distribution is polarised. And data on polarised radiation is

capable of identifying and distinguishing between synchrotron emission and inverse-Compton

emission as they in principle produce different degrees of polarisation. Thus, thus discovery

of polarised emission at high energies would be a test of models of blazar SEDs amongst

other things.

Now high energy emission could also be hadronic (e.g. Becker, 2008; Reimer, 2012; Doert

et al., 2012; Dermer et al., 2007). Hadronic processes involving protons produce high energy

radiation mainly gamma rays in the following ways. They first interact with either other

113

protons or photons (Becker, 2008).

p γ → ∆+ → p π0

→ n π+ (5.21)

p p→ π+ π− π0 (5.22)

The resulting pions decays produce gamma rays directly or indirectly. Neutral pions decay

to photons directly as

π0 → γ γ (5.23)

The charged pions decay into secondary leptons, that can then produce synchrotron or

inverse-Compton emission.

π+ → µ+ νµ → e+ νe νµ νµ (5.24)

Hadronic emission from blazars could also produce polarisation from the secondary lep-

tons. Once, again the leptons would produce IC scattering and synchrotron radiation. The

spectrum of this polarised radiation would be different from that produced by the primary

leptons. As a result, by spectro-polarimetric observations secondary leptonic gamma rays will

be identified and their relative proportion to other processes determined. Particularly as the

secondary leptonic emission can be related to its primary hadronic source through branching

ratios, it is possible to put indirect constraints on the primary hadronic emission via the lep-

tonic secondaries. And as, these leptons are again correlated by well defined branching ratios

to neutrinos, there are some prospects of a possible relation between neutrino and polarised

high energy emission, though it is likely to be model dependent and potentially complicated.

But these are exciting avenues that need to be explored in order to completely understand

those very complications that are so central to the emission mechanisms in blazars.

114

5.6 Discussion and conclusions

It is very clear that polarimetry is the next frontier in high energy astronomy that holds a

lot of promise. It is extremely challenging to perform polarimetric measurements in general.

Specifically for blazars, it is possible that instruments in space are required to make detec-

tions. However, in an ideal situation, i.e. if there are dedicated balloon flights performing

observations of previously selected sources that are flaring, then it is not inconceivable that a

signal maybe be observed. Also, remembering that high energy polarisation hasn’t yet been

measured and the predictions are based on theoretical assumptions albeit reasonable ones,

if the polarisation values happen to be as high as simulations in McNamara et al. (2009)

predict then balloon experiments will have an excellent chance of detecting something.

The strategy of using the low energy polarisation measurements of blazars from Fujiwara

et al. (2012); Blinov et al. (2013) and other similar work to select blazars looks promising

to effectively focus on the ones which are the brightest and most polarised. All of this is

expressed in terms of the MDP and required observation times. This is similar to the plan

proposed by the GEMS team Krawczynski et al. (2013). The numbers for GEMS-like mission

are consistent with those found by This is similar to the plan proposed by the GEMS team

Krawczynski et al. (2013). Of course, this study is not limited to GEMS like missions or that

energy range. Predictions have been made for a number of different polarimetric missions

such as PoGOLITE, Astro-H, POLARIX, GAP.

As discussed in section 5.4, flaring increases detectability of blazars. Monitoring flar-

ing activity of sources in optical and other wavelengths to determine flare statistics will

help build our database of promising candidates while waiting for upcoming polarimetry

missions. Low energy polarimetry missions like RoboPol can be used to guide high energy

polarimetry observations based on monitoring of flaring blazar as Blinov et al. (2013). And in

future, synergistic multiwavelength polarimetry observations will definitely increase chances

of discovering blazars in this new domain.

Polarimetry is critical to not just multiwavelength, but possibly multimessenger astro-

115

physics as pointed out in section 5.5. Neutrinos are indicators of primarily hadronic emission

mechanisms in blazars. Polarisation could be an indicator of both, though it is more directly

related to the primary leptonic emission. It is associated to hadronic mechanism via radia-

tive emission from the secondary electrons. Thus, polarimetry could probe both secondary

and primary emission and hadronic and leptonic emission. And more specifically, it can

distinguish between self-synchrotron and external Compton mechanism.

In addition, high polarimetry of blazars can potentially put constraints on Lorentz Invari-

ance Violation in much the same way as GRBs do. If blazars were to possess a reasonably

high degree of polarisation, being high redshift sources, they could be used for probing

LIV (e.g., Coleman & Glashow, 1999; Ellis et al., 2000; Laurent et al., 2011) using vacuum

birefringence.

Thus, high energy polarimetry is an exciting field with tremendous discovery potential.

There is a huge scope for theoretical work and observational and mission strategising in

preparation for future generation of polarimeters.

116

Chapter 6

Radio Supernovae in the GreatSurvey Era

Abstract

1 Radio properties of supernova outbursts remain poorly understood despite longstanding

campaigns following events discovered at other wavelengths. After ∼ 30 years of observa-

tions, only ∼ 50 supernovae have been detected at radio wavelengths, none of which are Type

Ia. Even the most radio-loud events are ∼ 104 fainter in the radio than in the optical; to date,

such intrinsically dim objects have only been visible in the very local universe. The detec-

tion and study of radio supernovae (RSNe) will be fundamentally altered and dramatically

improved as the next generation of radio telescopes comes online, including EVLA, ASKAP,

and MeerKAT, and culminating in the Square Kilometer Array (SKA); the latter should be

>∼ 50 times more sensitive than present facilities. SKA can repeatedly scan large (>∼ 1 deg2)

areas of the sky, and thus will discover RSNe and other transient sources in a new, auto-

matic, untargeted, and unbiased way. We estimate SKA will be able to detect core-collapse

RSNe out to redshift z ∼ 5, with an all-redshift rate ∼ 620 events yr−1 deg−2, assuming a

survey sensitivity of 50 nJy and radio lightcurves like those of SN 1993J. Hence SKA should

provide a complete core-collapse RSN sample that is sufficient for statistical studies of radio

properties of core-collapse supernovae. EVLA should find ∼ 160 events yr−1 deg−2 out to

redshift z ∼ 3, and other SKA precursors should have similar detection rates. We also pro-

vided recommendations of the survey strategy to maximize the RSN detections of SKA. This

new radio core-collapse supernovae sample will complement the detections from the optical

1This chapter matches the version which has been submitted to Astrophysical Journal, and posted versionon arxiv.org numbered on arXiv:1206.0770 and is co-authored with Brian Fields.

117

searches, such as the LSST, and together provide crucial information on massive star evolu-

tion, supernova physics, and the circumstellar medium, out to high redshift. Additionally,

SKA may yield the first radio Type Ia detection via follow-up of nearby events discovered

at other wavelengths.

6.1 Introduction

Supernovae are among the most energetic phenomena in the universe, and are central to

cosmology and astrophysics. For example, core-collapse supernovae are explosions arise

from the death of massive stars and hence are closely related to the cosmic star-formation

rate and to massive-star evolution, and are responsible for the energy and baryonic feedback

of the environment (Madau et al., 1998). Type Ia supernovae show a uniform properties in

their lightcurves and play a crucial role as cosmic “standardizable candles” (Phillips, 1993;

Riess et al., 1996).

Our knowledge of the optical properties of supernovae, is increasing rapidly with the

advent of prototype “synoptic”–i.e., repeated scan–sky surveys, such as SDSS-II (Frieman

et al., 2008; Sako et al., 2008) and SNLS (Bazin et al., 2009; Palanque-Delabrouille et al.,

2010). These campaigns are precursors to the coming “Great Survey” era in which synoptic

surveys will be conducted routinely over very large regions of sky, e.g., LSST 2 (LSST Science

Collaboration et al., 2009) and Pan-STARRS 3. The number of detected supernovae will

increase by several orders of magnitude in this era (LSST Science Collaboration et al., 2009;

Lien & Fields, 2009).

In contrast to this wealth of optical information, properties of supernovae in the radio

remain poorly understood, fundamentally due to observational limitations. Radio super-

novae (RSNe) have primarily been discovered by follow-up observations of optical outbursts,

and only very rarely by accident. To date, only ∼ 50 core-collapse outbursts have radio

2http://www.lsst.org/Science/docs/SRD.pdf3http://pan-starrs.ifa.hawaii.edu/project/science/precodr.html

118

detections, and no Type Ia explosion has ever been detected in the radio (Weiler et al., 2004;

Panagia et al., 2006). The core-collapse subtype Ibc has been a focus of recent study in the

radio, because some Type Ibc events are associated with long Gamma-Ray Bursts (GRBs)

(Galama et al., 1998; Kulkarni et al., 1998; Soderberg, 2007; Berger et al., 2003).

Current radio interferometers are scheduled primarily around targeted observations pro-

posed by individual principal investigators. This stands in contrast to future radio interfer-

ometers planned for the coming “Great Survey” era. These include the Square Kilometer

Array (SKA4) and its precursor prototype arrays (for example, ASKAP5 and MeerKAT6).

These telescopes will operate primarily as wide-field survey instruments focusing on several

key science projects (Carilli & Rawlings, 2004). As synoptic telescopes, they will be far bet-

ter suited to study all classes of transient and time-variable radio sources, including RSNe.

Gal-Yam et al. (2006) already pointed out the power of synoptic radio surveys for detecting

radio transients of various types, including supernovae and GRBs, in an unbiased way. Here

we quantify the prospects for RSNe.

In this paper we explore this fundamentally new mode of untargeted RSN discovery and

study. We adopt a forward-looking perspective, and consider the new science enabled by

RSNe observations in an era in which the full SKA is operational. Our focus is mainly on

core-collapse supernovae, the type for which some radio detections already exist. However,

we will also discuss the possibility of Type Ia radio discovery based on current detection

limits. We will first summarize current knowledge of radio core-collapse supernovae (§6.2),

and the expected sensitivity of SKA (§6.3). Using this information, we forecast the radio core-

collapse supernovae harvest of SKA (§6.4), and consider optimal survey strategies (§6.5). We

conclude by anticipating the RSN science payoff in this new era (§6.6). We adopt a standard

flat ΛCDM model with Ωm = 0.274, ΩΛ = 0.726, and H0 = 70.5 km s−1 Mpc−1 (Komatsu

et al., 2009) throughout.

4http://www.ska-telescope.org5http://www.atnf.csiro.au/projects/askap6http://www.ska.ac.za/meerkat

119

6.2 Radio Properties of Supernovae

Several key properties of RSNe have been established, as a result of the longstanding lead-

ership of the NRL-STScI group (recently reviewed in Weiler et al., 2009; Stockdale et al.,

2007; Panagia et al., 2006) and of the CfA group and others (summarized in Soderberg,

2007; Berger et al., 2003). We summarize these general RSN characteristics, which we will

use to forecast the RSN discovery potential of synoptic radio surveys.

6.2.1 Radio Core-Collapse Supernovae

Observed core-collapse RSNe have luminosities spanning νLν ∼ 1033 − 1038 erg s−1 at 5

GHz, and thus are >∼ 104 times less luminous in the radio than in the optical. Their intrinsic

faintness has prevented RSN detection in all but the most local universe. Even within a

particular core-collapse subtype, radio luminosities and lightcurves are highly diverse, e.g.,

two optically similar Type Ic events might be radio bright in one case and undetectable in

the other (Munari et al., 1998; Nakano & Aoki, 1997) 7 8. Additionally, core-collapse RSNe

spectral shapes strongly evolve with time; lightcurves peak over days to months depending on

the frequency. RSN emission can be understood in terms of interactions between the blast,

ambient relativistic electrons, and the circumstellar medium (Chevalier, 1982a,b, 1998).

To model RSN emission as a function of frequency and time, we adopt the semi-empirical

form derived by Chevalier (1982a) and extended in Weiler et al. (2002),

L(t, ν) = L1

( ν

5 GHz

)α(

t

1 day

e−τexternal

(

1 − e−τCSMclumps

τCSMclumps

) (

1 − e−τinternal

τinternal

)

. (6.1)

We follow the notation of Weiler et al. (2002). L(t, ν) is the supernova luminosity at fre-

quency ν and time t after the explosion. Optical depths from material both outside (τexternal,

τCSMclumps) and inside (τinternal) the blast-wave front are taken into account (see Weiler et al.,

7http://rsd-www.nrl.navy.mil/7213/weiler/sne-home.html8New Radio Supernova Results are available online at: http://rsdwww.nrl.navy.mil/7213/weiler/sne-

home.html

120

2002).

Parameters embedded in each optical depth term are those for SN 1993J, one of the best

studied RSNe (Weiler et al., 2007). Radio emission from SN 1993J is dominated by the

clumped-circumstellar-medium (clump-CSM) term, and hence

L(t, ν) ∼ 1 − e−τCSMclumps

τCSMclumps

, (6.2)

where τCSMclumps= 4.6 × 105 ( ν

5 GHz)−2.1 ( t

1 day)−2.83, for SN 1993J. At small t, τCSMclumps

is

large and L(t, ν) ∼ 1/τCSMclumps∝ ν2.1 t2.83, so luminosity grows as a power law at early

times. With all optical depth parameters fit to SN 1993J, the peak luminosity is controlled

by the prefactor L1.

Our main focus will be on RSN discovery, and thus it is most important to capture the

wide variety of peak radio luminosities, which correspond in our model to a broad distri-

bution for L1. Figure 6.1 shows a crude luminosity function (not-normalized) based on the

sample of 20 core-collapse supernovae (15 Type II and 5 Type Ibc) that have a published

peak luminosity at 5 GHz (Weiler et al., 2004; Stockdale et al., 2003, 2007; Papenkova et al.,

2001; Baklanov et al., 2005; Pooley et al., 2002). We use 5 GHz data to construct our lumi-

nosity function because the most RSNe have been observed at this frequency. However, our

predictions will span a range of frequencies, based on this luminosity function and eq. (6.2).

The data are divided into four luminosity bins of size ∆ log10(L) = 1. The black curve in

Fig. 6.1 is the best-fit Gaussian, with average luminosity log10(Lavg/erg s−1 Hz−1) = 27.3,

a standard deviation σ = 1.25, and χ2 = 0.18. SN 1987A is marked in Fig. 6.1, but was

not used in the fit to avoid possible bias due to its uncommonly low luminosity. The fitted

luminosity function might be biased towards the brighter end, because of the current survey

sensitivity and the small and incomplete nature of the sample.

121

Figure 6.1: Radio luminosity function (not-normalized) at 5 GHz of core-collapse supernovaeshowing core-collapse supernovae count as a function of log10(L), where L is the peak radioluminosity. Data are binned to ∆ log10(L) = 1. The black curve shows the χ2-fitted Gaussianto the underlying data (red stars).

6.2.2 Radio Type Ia Supernovae

All searches to date have failed to detect radio emission from Type Ia supernovae. Panagia

et al. (2006) reported the radio upper limits of 27 Type Ia supernovae from more than two

decades of observations by the Very Large Array (VLA). The weakest limit on a Type Ia

event is 4.2 × 1026 erg s−1 Hz−1 at 1.5 GHz for SN 1987N, which is around one order of

magnitude lower than the average luminosity of radio-detected core-collapse supernovae (see

§6.2.1). The strongest limit on Type Ia radio emission is even tighter, 8.1×1024 erg s−1 Hz−1

at 8.3 GHz for SN 1989B. Additionally, the z ∼ 0 cosmic Type Ia supernova rate is around

1/4.5 of the core-collapse supernova rate (Bazin et al., 2009). The intrinsic faintness in radio

and their smaller rate make detecting Type Ia in radio observations especially hard.

122

6.3 Next-Generation Radio Telescopes: Expected

Sensitivity

Radio detections of supernovae to date have been restricted both by the limiting sensitivity

of contemporary radio interferometers and the need for dedicated telescope time for transient

followup. This situation will change drastically with SKA’s unprecedented sensitivity and

particularly by its ability to repeatedly scan large regions of the sky at this great depth.

Current SKA specifications adopt a target sensitivity parameter Aeff/Tsys = 104 m2 K−1

at observing frequencies in the low several GHz, including z = 0 HI at 1.4 GHz. Aeff is the

effective aperture, and Tsys is the system temperature. We will adopt this value of Aeff/Tsys,

which yields a 1-σ rms thermal noise limit in total intensity of

σI = 0.15 µJy (∆ν/GHz)−1/2 (δt/hr)−1/2, (6.3)

for a bandwidth ∆ν and observation duration δt. The SKA will therefore reach a thermal

noise limit of several nJy in deep continuum integrations (δt ∼ 1000 hr) 9. We define the

associated survey sensitivity Smin (the minimum flux density threshold) as Smin = 3σI . In

common with other radio interferometers, SKA will accumulate sensitivity in targeted deep

fields, including transient-monitoring fields, by accumulating integration time over multiple

individual observing tracks. We therefore will adopt a fiducial SKA supernova sensitivity of

Smin = 50 nJy in 100 hours of observation, but we will show how our results are sensitive to

other choices of Smin.

It is anticipated that transient fields will be revisited with a cadence appropriate to

the variability timescales under study and that interferometric inverse imaging methods will

include source models with time variability. Survey optimization for interferometric transient

detection is an active area of current SKA research. The technical details are beyond the

intent and scope of this paper, but will be influenced by science goals for transient source

9http://www.skatelescope.org/PDF/DRM v1.0.pdf

123

study in general, including pulsars, GRBs, and supernovae (as considered in this paper), as

well as the as-yet undiscovered transient population.

6.4 Radio Supernovae for SKA

With its unprecedented sensitivity, SKA will be capable of synoptic search for core-collapse

RSNe and open new possibilities in radio astronomy. In this section, we predict the RSN

detections of SKA based on current knowledge to demonstrate how the RSN survey can be

done.

6.4.1 Core-Collapse Supernovae

The detection rate Γdetect = dNSN/(dtobs dΩ dz) for a given RSN survey is

Γdetect = fsurvey fradio fISM ΓSN , (6.4)

and is set by several observability factors f that modulate the total rate of all supernovae

ΓSN(z) =dNSN

dVcomov dtem

dtemdtobs

dVcomov

dΩ dz= RSN(z) r2

comov(z) c

dt

dz

(6.5)

within the cosmic volume out to redshift z (Madau et al., 1998; Lien & Fields, 2009).

We see that the total cosmic supernova rate ΓSN depends on cosmology via the volume el-

ement and the time dilation terms. Because ΛCDM cosmological parameters are now known

to high precision, these factors have a negligible error compared to the other ingredients in

the calculation. The other factor in ΓSN is the cosmic core-collapse supernova rate density

RSN(z) = dNSN/(dVcomov dtemit). Some direct measurements of this rate now exist out to

z ∼ 1, but the uncertainties remain large (Cappellaro et al., 1999; Dahlen et al., 2004; Cap-

pellaro et al., 2005; Hopkins & Beacom, 2006; Botticella et al., 2008; Dahlen et al., 2008;

Kistler et al., 2011; Bazin et al., 2009; Smartt et al., 2009; Dahlen et al., 2010; Li et al.,

124

2011c; Horiuchi et al., 2011). However, core-collapse events are short-lived, and so the cos-

mic core-collapse rate is proportional to the cosmic star-formation rate ρ⋆, which is much

better-determined and extends to much higher redshifts. We thus derive RSN from the re-

cent Horiuchi et al. (2009b) fit to the cosmic star-formation rate. The proportionality follows

from the choice of initial mass function; we apply the Salpeter initial mass function (Salpeter,

1955) and assume the mass range of core-collapse SNe progenitors to be 8M⊙ − 50M⊙; this

gives RSN = (0.007M−1⊙ ) ρ⋆

Several effects reduce the total rate ΓSN to the observed rate Γdetect in eq. (6.4). Due to

finite survey sensitivity, only a fraction fsurvey of events are bright enough to detect, and only

some fraction fradio of supernovae will emit in the radio. We neglect interstellar extinction

and assume fISM ∼ 1 at the radio wavelengths considered.

The term fradio in eq. (6.4) contains the greatest uncertainty due to the relatively small

sample of RSNe observed to date, and the unavoidable incompleteness of the sample (K.

Weiler, private communication 2010). The only published fraction available is for Type Ibc

supernovae. Using VLA for radio follow-up, Berger et al. (2003) suggests that fradio,Ibc ∼ 12%

after surveying 33 optically-detected Type Ibc supernovae. For the purpose of demonstration,

we will adopt fradio = 10% for the calculations presented in this paper, which we believe is

rather conservative.

An order-of-magnitude calculation provides a useful estimate of the expected core-collapse

RSN rate. As discussed in §6.3, we adopt a fiducial SKA sensitivity of Smin = 50 nJy. Hence

SKA will be able to detect supernovae with average radio luminosity (L ∼ 1027 erg s−1 Hz−1)

to a distance DL =√

L/4πSmin ∼ 4 Gpc, which for a ΛCDM cosmology corresponds to

z ∼ 1. This will give a detectable volume of Vdetect ∼ (4/3)πD3L ∼ 2.85 × 1011 Mpc3.

Observations show that the core-collapse supernova rate RSN ∼ 10−3 yr−1 Mpc−3 at z ∼ 1

(Dahlen et al., 2008, 2010). Assuming the fraction of the total core-collapse supernovae that

display the adopted average radio luminosity to be fradio ∼ 10% (Berger et al., 2003), the

all-sky detection rate dNSN/dt ∼ RSN × fradio × Vdetect ∼ 2.85 × 107 yr−1. This corresponds

125

to a areal detection rate dNSN/(dt dΩ) ∼ 700 yr−1 deg−2. As we now see, a more careful

calculation confirms this estimate.

Figure 6.2: Estimated radio core-collapse supernova detection rate as a function of redshiftat 1.4 GHz, assuming fradio = 10%. Predictions are shown for different survey sensitivities:Smin = 10 µJy (blue), 1 µJy (blue), 100 nJy (blue), 50 nJy (thick red), 10 nJy (blue), 1nJy (blue) from bottom to top solid curves, respectively. We adopt 50 nJy as our bench-mark sensitivity hereafter. For comparison, the red-dashed curve shows the LSST opticalsupernova detection rate per year per deg2 (Lien & Fields, 2009). Also, the top solid curve(black) plots the ideal core-collapse RSN rate for comparison.

A careful prediction involves detailed calculation of fsurvey(z). The fraction fsurvey(z) of

observable radio-emitting events depends on adopted survey sensitivity, and on the normal-

ized supernova luminosity function Φ5GHz(logL), which is measured at a peak luminosity at

5 GHz (derived in § 6.2). In this paper we will only consider whether a supernova is de-

tectable at its peak luminosity at each corresponding frequency. The peak radio luminosity

should be reached earlier at higher frequencies because of preferential absorption at lower

frequencies (Weiler et al., 2002). At different redshift, the peak flux density Speakmin in the

126

Figure 6.3: Estimated radio core-collapse supernova detection rate as a function of redshiftfor different frequency bands, for fradio = 10%, and an adopted survey sensitivity Smin = 50nJy.

Figure 6.4: Core-collapse detection sensitivity to supernova radio luminosity, at 1.4 GHz,and for survey sensitivity Smin = 50 nJy. (a)Left Panel: Supernova distribution over redshift,for different cutoffs for the luminosity function. (b)Right Panel: Supernova distribution inluminosity bins, integrated over all redshifts.

observed frequency ν can be tied to the corresponding luminosity threshold Lpeakmin by

Lpeakmin (z; νemit) =

4πD2L(z)

(1 + z)Speak

min (νobs), (6.6)

127

where the luminosity distance isDL(z) = (1+z) c/H0

∫ z

0dz′ [Ωm(1+z′)3+ΩΛ]−1/2 . However,

because the luminosity function we used is based on the peak luminosity at 5 GHz, we must

find the corresponding luminosity threshold at this frequency by applying corrections based

on the radio spectrum,

Lpeakmin,5GHz = Lpeak

min

5GHz bandSpeak(νem) dνem

obs bandSpeak[(1 + z)νobs] dνobs

. (6.7)

The detectable fraction resulting from survey sensitivity can therefore be estimated as

fsurvey =

log Lpeakmin,5GHz

Φ5GHz(logL) d logL. (6.8)

Figure. 6.2 plots the results of our predicted core-collapse RSN detection rate for different

target survey sensitivities, Smin. We adopt a benchmark frequency of 1.4 GHz because this

will be one of the first major bands SKA deploys to observe neutral hydrogen. The related

instantaneous field-of-view at 1.4 GHz of current SKA designs based on dish reflectors is

approximately 1 deg2, which we adopt. Fig. 6.2 plots the ideal core-collapse supernova

rate for comparison (assuming infinite sensitivity but fradio = 10%). One can see that the

detection rate at 1 nJy is very close to the ideal rate in the universe.

Results for our fiducial SKA flux limit Smin = 50 nJy are highlighted in Fig. 6.2. At this

sensitivity, we see that we can expect that radio supernovae will be discovered (event rates

> 5 RSN yr−1 deg−2) over the enormous redshift range z ≃ 0.5 to 5. The total rate of RSNe

expected in this entire redshift range is

dNSN

dt dΩ(> 50 nJy) ≈ 620 RSNe yr−1 deg−2 , (6.9)

in good agreement with our above order-of-magnitude estimate. This sample size is large

enough to be statistically useful and to allow for examination of the redshift history of RSNe.

Moreover, out to z ∼ 1, SKA will detect almost all cosmic RSNe in the field of view, while

128

at higher redshift the detections still comprise > 10% of the underlying ideal cosmic rate.

For comparison, we also see that LSST will detect optical supernovae out to z <∼ 1. Thus

SKA will be complementary to LSST as a unique tool for cosmic supernova discovery.

Figure. 6.3 shows how core-collapse RSN detections vary for different observing frequen-

cies, fixing a common survey sensitivity Smin = 50 nJy and bandwidth ∆ν = 1 GHz. Results

show similar numbers of detections at different bands, which is caused by a relatively flat

spectrum shape at peak luminosities. Because SKA will be able to detect core-collapse RSNe

out to high redshift z ∼ 5, the frequency-redshift and time-dilation effects are significant.

Weiler et al. (2002) noted that RSNe peak when the optical depth τ ∼ 1. Since the optical

depth depends both on frequency and time with similar power index (Weiler et al., 2002), the

frequency-redshift and time-dilation effects approximately cancel, so that a fixed observed

frequency, the peak time is nearly redshift-independent.

As mentioned above, our luminosity function is likely biased toward the available bright

events in a small and incomplete sample. To explore how this bias could affect our results,

Fig. 6.4 shows how the detection rate with Smin = 50 nJy at 1.4 GHz depends on core-

collapse RSN luminosity. Fig. 6.4(a) shows that RSN with peak luminosities greater than

1027 erg s−1 Hz−1 contribute all of the detections beyond redshift z ∼ 1, and RSN need to

peak brighter than 1028 erg s−1 Hz−1 to be seen beyond z ∼ 3. Fig. 6.4(b) similarly shows that

the all-redshift detection rate becomes substantial for explosions peaking > 1026 erg s−1 Hz−1.

Type Ibc supernovae are of particular interest given their association with long gamma-

ray bursts (GRBs; Galama et al., 1998; Woosley, 1993; Heger et al., 2003). Fig. 6.5 shows

our predictions for Type Ibc detections of SKA per year per deg2 at 1.4 GHz with a survey

sensitivity of 50 nJy. The red curve shows the radio Type Ibc detections, assuming that

Type Ibc represents 25% of core-collapse events (Li et al., 2011b), and fradio,Ibc = 12% with

luminosity ∼ 1027 erg s−1 Hz−1 10 (Berger et al., 2003). The blue curve shows the possi-

ble detections of the sub-class of Type Ibc supernovae that display extreme radio emission

10Here we simply assume a Gaussian distribution for the luminosity function centered at the specifiedluminosity with σ = 1.

129

and hence might be powered by central engines and related to GRBs. We assume that

0.5% of all Type Ibc supernovae are powered by central engines and have luminosities of

∼ 1029 erg s−1 Hz−1 (Berger et al., 2003). We adopted the spectrum of SN 1998bw, which

is a Type Ic supernova (Weiler et al., 2002). Under these assumptions the SKA will be able

to make unbiased, untargeted detections of ∼ 130 radio Type Ibc supernovae per year per

deg2, and ∼ 20 Type Ibc supernovae that might be connected to GRBs.

Figure 6.5: Predicted detection rate of Type Ibc supernovae as a function of redshift. Inthis plot we assume the sensitivity for SKA is Smin = 50 nJy. The red curve shows all of theradio Ibc detections, assuming fradio,Ibc = 12% (Berger et al., 2003). The blue curve showsonly the detection rate for Radio Ibc with central engines, assuming 0.5% of all of the TypeIbc RSNe are powered by central engines.

Finally, we turn to SKA precursors. The EVLA11, a current leading-edge radio inter-

ferometer operating at centimeter wavelengths, is anticipated to reach a 1-σ rms noise of

σI ∼ 1 µJy or less in 10 hours, while SKA is expected to reach σI ∼ 50 nJy in 10 hours.

With data accumulated over repeated scans spanning over 1000 hours, an rms σI ∼ 5 nJy

may be reached. In synoptic surveys, we would expect EVLA to detect core-collapse events

11http://www.aoc.nrao.edu/evla

130

at a rate ∼ 160 RSNe yr−1 deg−2 over a redshift range z = 0.5 to 3 (Fig. 6.2). A sample of

this size over this redshift range will already mark a major advance in the study of cosmic

RSNe, and further motivate the full SKA. ASKAP and MeerKAT are expected to have sen-

sitivities comparable to that of EVLA (Johnston et al., 2009; de Blok et al., 2010), hence we

would expect these to detect RSNe with similar rates and redshift reach.

6.4.2 Type Ia Supernovae

If all Type Ia RSNe are dimmer than the weakest limit presented in §6.2.2, the expected SKA

detection rate is essentially zero. For example, if a typical Type Ia has a radio luminosity

equal to the lowest published limit, L = 8.1 × 1024 erg s−1 Hz−1, this can be seen with a

sensitivity Smin = 50 nJy out to a luminosity distance ∼ 300 Mpc (z ∼ 0.08). While ∼ 3900

cosmic Ia events should occur per year out to this distance over the entire sky, ≪ 1 events

are expected in the SKA field of view. More optimistically, imagine a typical Type Ia radio

luminosity is L = 1026 erg s−1 Hz−1, which is below L = 4.2× 1026 erg s−1 Hz−1, the highest

published limit (Panagia et al., 2006); here the luminosity distance increases to ∼ 1400 Mpc

(z ∼ 0.28). In this case, we find an SKA Type Ia detection rate ∼ 0.5 yr−1 deg−2, based

on the local cosmic Type Ia rate derived from SDSS-II optical data (Dilday et al., 2010),

Smin = 50 nJy, and fradio = 10%.12 We see that even optimistically, we expect fewer than

one event per SKA field-of-view per year. Even with fradio = 100%, the detection rate is still

only ∼ 5 yr−1 deg−2. Therefore we conclude that SKA will make few, if any, blind detections

of Type Ia supernovae.

Targeted radio observations to follow up from nearby optical detections will probably

be the best way to search for such events. For example, we expect 10 Type Ia events/year

in the LSST sky within ∼ 60 Mpc (z ∼ 0.015). Type Ia (or core collapse!) events within

this distance observed with Smin = 50 nJy, would be detectable at luminosities L >∼ 3.0 ×

1023 erg s−1 Hz−1. Amusingly, this is close to the radio luminosity of SN 1987A.

12This also is implied by Fig. 6.4, which is for core-collapse events that have a higher cosmic rate.

131

6.5 Radio Survey Recommendations

A key requirement for detecting weak radio emission from CSM-supernovae interactions is

improved radio interferometer sensitivity. High angular resolution – below an arcsecond

at 1.4 GHz, (Weiler et al., 2004) – is also required to avoid natural confusion and to help

identify supernovae against background galaxies. This is similar to the maximum EVLA

angular resolution at 1.4 GHz. For comparison, the maximum anticipated SKA baseline

length of 3000 km, producing angular resolution of 0.014 arcsecond at 1.4 GHz, is sufficient

to distinguish different galaxies and also to resolve galaxies as extended sources within the

observable universe with rms confusion limit of < 3 nJy at 1.4 GHz (Carilli & Rawlings,

2004).

A key science goal of the SKA is to detect transient radio sources, both known (e.g.

pulsars, GRBs), and as-yet unknown. This requires sophisticated transient detection and

classification algorithms very likely running commensally with other large surveys planned

by the SKA, such as the HI spectroscopic survey and deep continuum fields. We assume

here that SKA transient detection algorithms will encompass automated detection of RSNe.

For example, current parameterized models (Weiler et al., 1986, 1990; Montes et al., 1997;

Chevalier, 1982a,b) based on available data predict patterns of spectral index evolution char-

acteristic of supernovae in general, and supernova sub-types in particular. This information

could be exploited for RSN detection, even potentially against a background of unrelated

source variability. Broad frequency coverage is important in this regard (Weiler et al., 2004).

The SKA intrinsically is a high dynamic-range instrument, given the sensitivity implied

by the large collecting area. The most demanding SKA science applications will require a

dynamic range of 107:1. The detection of faint RSNe will require a dynamic range that falls

within that envelope.

Although the lightcurves of RSNe show great diversity, the luminosities of core-collapse

supernovae usually change much slower in radio than in optical. RSN lightcurves typically

evolve on timescales of weeks to years; a useful lightcurve compilation appears in Stockdale

132

et al. (2007). Thus the minimum survey cadence (revisit periodicity) need not be any more

frequent than this. Also, we have shown that core-collapse RSNe can be found out to high

redshift with surveys pushing down to Smin = 50 nJy. For SKA this corresponds to about

100 hours of exposure, in line with planned deep field exposures which are part of the key

science. Thus, SKA as currently envisioned is well-suited to core-collapse discovery.

On the other hand, SKA probably will not have sufficient survey sensitivity for a volu-

metric search for Type Ia events, based on our current knowledge of the cosmic Type Ia rate

and the upper limits in their luminosities set by the non-detection of these events. Follow-up

observations from other wavelengths will likely be the best way to search for Type Ia RSNe.

The small volume of the local universe will limit nearby untargeted SKA detections of

low-redshift core-collapse RSNe. We estimate only ∼ 2 core-collapse RSN detections per

year per square degree within redshift z ∼ 0.5 (assuming a 50 nJy sensitivity at 1.4 GHz

and fradio = 10%). Unless SKA has large sky coverage comparable to those of optical

surveys, it will be hard to get statistical information from such a small sample. Therefore,

targeted radio followup of optically-confirmed nearby supernovae will be crucial to build a

core-collapse RSN “training set” database needed for refining automatic identification and

classification techniques.

With detection methods optimized based on low-z radio data for optically-identified

events, radio surveys can then be used to independently detect core-collapse RSNe at high

redshift based only on their radio emission. As shown in Fig. 6.2, supernova searches at

high redshift (z & 1) will largely rely on radio synoptic surveys, the inverse of the strategy

proposed above for low-redshift domain. Surveys for core-collapse RSNe will likely not be

synoptic all-sky surveys due to operational limitations, but will likely proceed in a limited

set of sub-fields, visited over an hierarchical set of cadences to cover a range of time-scales for

general transient phenomena and multiple commensal science objectives. It is also important

to match core-collapse RSNe survey sky coverage and cadence to that used in complementary

optical surveys. LSST will repeatedly scan the whole sky every ∼ 3 days. Thus a cadence

133

∼ 1 week for RSNe sub-fields will be preferred for an SKA core-collapse supernova survey.

6.6 Discussion and Conclusions

SKA’s capability for unbiased synoptic searches over large fields of view will revolutionize the

discovery of radio transients in general and core-collapse RSNe in particular (Gal-Yam et al.,

2006). The unprecedented sensitivity of SKA could allow detection of core-collapse RSNe out

to a redshift z ∼ 5. These detections will be unbiased and automatic in that they can occur

anywhere in the large SKA field of view without need for targeting based on prior detection

at other wavelengths. With SKA, the core-collapse RSN inventory should increase from the

current number of several dozen to ∼ 620 yr−1 deg−2. EVLA should detect ∼ 160 yr−1 deg−2,

and other SKA precursors should reap similarly large RSN harvests. In contrast, intrinsically

dim RSNe such as Type Ia events and 1987A-like core-collapse explosions are unlikely to be

found blindly. However, the SKA (and precursor) sensitivities will offer the possibility of

detecting these events via targeted followup of discoveries by optical synoptic surveys such

as LSST.

The science payoff of large-scale RSNe searches touches many areas of astrophysics and

cosmology. We conclude with examples of possible science applications with the new era of

RSN survey. However, the true potential of untargeted radio search is very likely beyond

what we mention.

Non-prompt RSN emission requires the presence of circumstellar matter, so such sur-

veys will probe this material and the pre-supernova winds producing it. For core-collapse

supernovae, pre-supernova winds should depend on the metallicities of the progenitor stars

(Leitherer et al., 1992; Kudritzki & Puls, 2000; Vink et al., 2001; Mokiem et al., 2007), and

should be weaker in metal-poor environments with lower opacities in progenitor atmospheres.

This effect should lead to correlations between RSN luminosity and host metallicity, as well

as an evolution of the RSN luminosity function towards lower values at higher redshifts. For

134

Type Ia supernovae, the mass-loss rate from the progenitors depends on the nature of the

binary system, i.e., single or double degenerate (Nomoto et al., 1984; Iben & Tutukov, 1984;

Webbink, 1984). Radio detection of Type Ia supernovae will probe the mass and density

profile of the surrounding environment and hence be valuable for studying Type Ia physics

(Eck et al., 1995; Panagia et al., 2006; Chomiuk et al., 2011).

Large-scale synoptic RSNe surveys will complement their optical counterparts. While op-

tical surveys such as LSST will provide very large supernova statistics at z . 1, radio surveys

will be crucial for detections at higher redshifts. The nature and evolution of dust obscu-

ration presents a major challenge for optical supernova surveys and supernova cosmology.

Current studies suggest dust obscuration increases rapidly with redshift, but uncertainties

are large. Mannucci et al. (2007) estimate that optical surveys may miss ∼ 60% of core-

collapse supernovae and ∼ 35% of Type Ia supernovae at redshift z ∼ 2. RSN observations,

in contrast, are essentially unaffected by dust. Thus, high-redshift supernovae could be de-

tected at radio wavelengths but largely missed in counterpart optical searches. Comparing

supernova detections in both optical and radio will provide a new and independent way to

measure dust dependence on redshift. In particular, SKA will be a powerful tool to directly

detect supernovae in dust-obscured regions at large redshift, and therefore offer what may

be the only means to study the total supernovae rate, star-formation, and dust behavior in

these areas.

Additionally, radio surveys will reveal rare and exotic events. For example, some Type

Ibc supernovae are linked to long GRBs (Galama et al. 1998; and see reviews in Woosley &

Bloom, 2006; Gehrels et al., 2009), probably via highly relativistic jets powered by central

engines and will manifest themselves in extremely luminous radio emission (Woosley, 1993;

Iwamoto et al., 1998; Li & Chevalier, 1999; Woosley et al., 1999; Heger et al., 2003). Thus

one might expect radio surveys to preferentially detect more Type Ibc supernovae than

other supernova types. An unbiased sample of Type Ibc RSNe will provide new information

about the circumstellar environments of these explosions and thus probe the mass-loss effects

135

believed to be crucial to the Ibc pre-explosion evolutionary path (Price et al., 2002; Soderberg

et al., 2004, 2006a; Crockett et al., 2007; Wellons & Soderberg, 2011); in addition, a large

sample of Ibc RSNe will allow systematic study of the differences, if any, between those which

do an do not host GRBs (Berger et al., 2003; Soderberg et al., 2006b; Soderberg, 2007).

Furthermore, radio surveys give unique new insight into a possible class of massive star

deaths via direct collapse into black holes, with powerful neutrino bursts but no electromag-

netic emission (MacFadyen & Woosley, 1999; Fryer, 1999; MacFadyen et al., 2001; Heger

et al., 2003). These “invisible collapses” can be probed by comparing supernovae detected

electromagnetically and the diffuse background of cosmic supernova neutrinos (Lien et al.,

2010, and references therein). By revealing dust-enshrouded SNe, radio surveys will make

this comparison robust by removing the degeneracy between truly invisible events and those

which are simply optically obscured. Indeed, direct collapse events without explosions but

with relativistic jets are candidates for GRB progenitors. A comparison among RSNe, opti-

cal supernovae, GRBs, and neutrino observations will provide important clues to the physics

of visible and invisible collapses, and their relation with GRBs.

We thus believe that a synoptic survey in radio wavelengths will be crucial in many

fields of astrophysics, for it will bring the first complete and unbiased RSN sample and

systematically explore exotic radio transients. SKA will be capable of performing such an

untargeted survey with its unprecedented sensitivity. Our knowledge of supernovae will thus

be firmly extended into the radio and to high redshifts.

6.7 Acknowledgments

We thank Kurt Weiler, Christopher Stockdale, and Shunsaku Horiuchi for encouragement

and illuminating conversations. We are also grateful to Joseph Lazio for helpful comments

that have improved this paper.

136

Chapter 7

Conclusions and Future Work

7.1 Discussion and Conclusions

The Standard Model “Foregrounds” in nuclear and particle astrophysics are interesting by

themselves as sites of interesting and often extreme astrophysics. In addition, it is critical

to understand them in order to be able to disentangle them from signatures of new physics

and astrophysics. The lithium problem is a classic case in point in the early universe. The

effect of standard nuclear physics needs to be disentangled from more exotic particle physics

effects. The search for nuclear resonances that can destroy 7Li or 7Be effectively represents

an effort to do this. The production reactions are very well understood leaving no room for

any addition or modification to the production network. So the destruction reactions are

the only options. The result of the study revealed that despite the large number of apriori

possible resonance channels involving n,p, d, t, 3He, 4He bombarding and destroying 7Li

and 7Be nuclei, very few channels have rates high enough to bring the abundances of 7Li

down to observed values. After systematically scanning the space of resonance strengths

and energies, only three candidate resonances remain and are enlisted once again here in the

table 7.1.

This work along with few others (For e.g., Cyburt & Pospelov, 2012; Boyd et al., 2010a),

prompted experiments (Kirsebom & Davids, 2011; O’Malley et al., 2011; Charity et al., 2011)

to look for these resonances. These experiments have essentially ruled out the possibility

of these resonances solving the lithium problem (Coc, 2013). This strengthens the exciting

possibility of the lithium discrepancy being an indication of new physics beyond the Standard

137

Compound Nucleus, Initial Linit Lfin Eres Γtot Exit Exit ChannelJπ, Eex State Channels Width

9B, (5/2+), 16.71 MeV 7Be + d 1 0 219.9 keV unknown p+ 8Be∗

(16.63 MeV) unknown1 α + 5Li unknown

10B, 7Be + t 1 1 130.9 keV < 600 keV p+ 9Be∗

(11.81 MeV) unknown2+, 18.80 MeV 1 3He unknown

2 α unknown10C, 7Be + 3He unknown unknown unknown unknown p unknown

unknown unknown (Q = 15.003 MeV) α unknown3He (elastic) unknown

Table 7.1: This table lists surviving candidate resonances.

Model. Of course, the possibility of a refined interpretation of the observational data to

reconcile with theoretical predictions, in terms of improved models of stellar atmospheres is

also very promising.

The study of properties of cosmic gamma-ray sources has been a very interesting one.

The computation of the inverse-Compton emission from star forming galaxies shows that it is

subdominant compared to pionic emission in most of the Fermi -LAT range of energies, unless

starburst galaxies have a much higher e/p ratio compared to normal galaxies, suggested by

some groups (Lacki et al., 2011). Aside from this caveat, it appears that the gamma-ray

emission from star forming galaxies is dominated by the pionic emission. This is in the pure

luminosity evolution model where only the luminosity of the galaxies evolves with redshift

and not the number density. In the pure density evolution case, the pionic emission is nearly

4 times weaker than the pure luminosity evolution for normal galaxies (Fields et al., 2010).

The linear scaling of the inverse-Compton emission with the star formation rate, implies

that the IC result remains the same as in the pure luminosity evolution case. And hence

for normal galaxies, the pionic emission will dominate in either case. For starbursts this is

not as obvious. According to high e/p models (Lacki et al., 2011) consistent with low values

of magnetic fields (50µG instead of 300µG), the IC emission can dominate over the pionic

emission at Fermi -LAT energies.

From the overall gamma-ray intensity of star forming galaxies as a function of energy, they

are a guaranteed and potentially significant contribution. And this is a subject of ongoing

138

research. However, it leaves the possibility of other sources as blazars also contributing

competitively (Stecker & Venters, 2011), particularly at high energies. This means that it

is important to have other “tie-breakers” to answer the question of relative contribution

of these sources to the diffuse EGB. This provides strong motivation to look at photon

statistics of these sources as a possible tie-breaker. From the preliminary results, it seems

inevitable that a EGB has a uniformly, multicomponent origin. That is both blazars and

star forming galaxies contribute without either one being dominant. This includes photons

from resolved as well as unresolved sources. The star-forming galaxies being numerous but

dim produce a largely Poisson distribution of photon counts. This is of course, contaminated

by an irreducible Poisson emission from the Galaxy. Thus, the normalisation of the Poisson

component represents an upper limit to the contribution of star forming galaxies. The blazars

on the other hand being brighter but less numerous, are capable of producing high count

pixels. As a result, their photon distribution has a high count tail. This is a distinguishing

signature between star forming galaxies and blazars. Preliminary results indicate that blazars

contribute to 21% of all the photons consistent with other analyses (Malyshev & Hogg, 2011;

Abdo et al., 2010l) including the LAT team’s own analysis.

Gamma-ray and X-ray polarimetry represents the next frontier in high energy astronomy.

The objects of interest here in this research and sources of cosmic rays in general are promis-

ing targets for future generations of polarimeters. Blazars and GRBs are in particular the

most exciting ones. This is because, they are both touted as candidates for producing ultra

high energy cosmic rays. And in general, they are regarded as the most powerful particle

accelerators in the universe. The accelerator properties are intimately tied to the high energy

emission mechanisms probed by polarimetry. Despite the fact that the polarimeters are at

most in the soft gamma-ray energy range, the relations between different energy bands from

energy exchanges through various radiative processes obeying energy conservation (Becker,

2008), implies that polarimetry can potentially constrain cosmic ray related emission mech-

anisms. Polarised emission from GRBs has been measured by SPI on INTEGRAL (For eg

139

Gotz et al., 2009), albeit for very few sources. Blazars on the other hand may require much

more sensitive space based polarimeters to be detected in this domain. Furthermore, po-

larised emission should largely trace the leptonic emission as inverse-Compton scattering of

low energy photons by electrons and synchrotron emission are likely to be the most impor-

tant mechanisms producing polarised X-rays and soft gamma-rays. And both synchrotron

and IC with different degrees of polarisation, can be distinguished by high energy polarisa-

tion. Hadronic emission on the other hand, dont lead to polarised emission for the processes

involving primary particles or reactions. The secondary electrons produced from hadronic

processes are however capable of producing polarised emission like for the primary leptons,

however with a different spectrum. Through the secondary leptonic emission polarised emis-

sion can indirectly probe hadronic processes. It could also therefore be potentially related

to neutrinos.

7.2 Future work

Naturally one of the promising future directions is the science of high energy polarimetry.

Observationally, estimating what fraction of blazars and GRBs would be discovered by fu-

ture generations of polarimeters, is one obvious goal. This involves doing feasibility studies,

optimising detection strategies, and hopefully eventually proposing for actual observations.

Theoretically, studying the emission mechanisms of both blazars and GRBs is a major chal-

lenge primarily in details of the various competing models. But chosing models that can

make predictions testable by future polarisation missions would be one way to approach this

task. Incorporating the multimessenger approach by connecting to neutrinos is an exciting

avenue.

I will join the HESS and CTA collaborations. This provides an opportunity to extend

my gamma-ray experience to very high energies. Once again these are testing grounds for

some of those very objects that have been of interest to me during my thesis. Cosmic

140

accelerators like star forming galaxies, blazars, GRBs, etc. are all very interesting at TeV

energies. Properties of individual sources like spectra of galaxies both active and otherwise,

flaring of blazars, etc. are areas of exciting research with the improved sensitivity of HESS

II. Furthermore, the possibility of testing the line emission from dark matter with HESS II

is extremely exciting.

Thus, I wish to continue to explore particle astrophysics with newer tools and methods.

141

Appendix A

A.1 The Narrow Resonance Approximation

Consider a reaction A + b → C∗ → c + D, which passes through an excited state of the

compound nucleus C∗. We treat separately normal and subthreshold reactions, defined

respectively by a positive and negative sign of the resonance energy Eres = Eex −QC , where

Eex is the excitation energy of the C∗ state considered, and QC = ∆(A) + ∆(B) − ∆(C∗).

In general, the thermally averaged rate is

〈σv〉 =

d3v e−µv2/2Tσv∫

d3v e−µv2/2T=

8

πµT−3/2

∫ ∞

0

dE E σ(E)e−E/T (A.1)

For a Breit-Wigner resonance with widths not strongly varying with energy, this becomes

〈σv〉 =4πωΓinitΓfin

(2πµT )3/2

∫ ∞

0

dEe−E/T

(E − Eres)2 + (Γtot/2)2(A.2)

Thus the thermal rate is controlled by the integral of the Lorentzian resonance profile mod-

ulated with the exponential Boltzmann factor.

The narrow resonance approximation has usually only been applied to the normal reso-

nance case, and assumes that the total resonance width is small compared to the temperature:

Γtot ≪ T .

142

A.1.1 Narrow Normal Resonances

In the normal or “superthreshold” case, the integral includes the peak of the Lorentzian

where E = Eres. The narrow condition then guarantees that over the Lorentzian width, the

Boltzmann factor does not change appreciably, and so we make the approximation

exp

(

−ET

)

≈ exp

(

−ET

)

(A.3)

where we choose the “typical” energy to be the peak of the Lorentzian, E = Eres. Then the

integral becomes

〈σv〉Γtot≪T ≈ ωΓinitΓfin

2(2πµT )3/2e−Eres/T

∫ ∞

0

dE1

(E − Eres)2 + (Γtot/2)2(A.4)

Furthermore, it is usually also implicitly assumed that the resonance energy is large compared

to the width: Eres ≫ Γtot. Then the integral gives 2π/Γtot, and the thermally averaged cross-

section under this approximation is given by Angulo et al. (1999),

〈σv〉Γtot≪T,Eres = ω Γeff

(

µT

)3/2

e−Eres/T (A.5)

= 2.65 × 10−13µ−3/2 ω Γeff T−3/29 exp(−11.605 Eres/T9) cm3s−1 (A.6)

where the latter expression has T9 = T/109 K.

Note however, that eq. (A.2) is exactly integrable as it stands and does not require we

make the usual Eres ≫ Γtot approximation. Thus for the normal case we modify the usual

reaction rate and instead adopt the form

〈σv〉narrow,normal = 〈σv〉Γtot≪T,Eres f(2Eres/Γtot) (A.7)

Here we introduce a temperature-independent correction for finite Eres/Γtot (still with Eres >

143

0)

f(u) =1

2+

1

πarctan u . (A.8)

This factor spans f → 1/2 for Eres ≪ Γtot to f → 1 for Eres ≫ Γtot.

In practice, we adopt a slightly modified version of the correction factor in our plots.

Recall that in Figs. 2.2–2.9, we show results for lithium abundances in the presence of

resonant reactions with fixed input channels, but without reference to a specific final state.

Without the correction factor, the resonant reaction rate is characterized by two parameters,

Eres and Γeff . These two parameters are insufficient to specify the correction factor, which

depends on Eres/Γtot. Rather than separately introduce Γtot, we instead approximate the

correction factor as f(2Eres/Γeff). Because Γeff < Γtot and f is monotonically increasing, this

always underestimates the value of f and thus conservatively understates the importance of

the resonance we seek (but the approximation is never off by more than a factor of 2 in the

normal case).

A.1.2 Narrow Subhreshold Resonances

Still making the narrow resonance approximation Γtot ≪ T , we now turn to the subthreshold

case, in which Eres < 0. To make effect of the sign change explicit, we rewrite eq. (A.2) as

〈σv〉 =ωΓinitΓfin

2(2πµT )3/2

∫ ∞

0

dEe−E/T

(E + |Eres|)2 + (Γtot/2)2(A.9)

Now the integrand always excludes the resonant peak, and only includes the high-energy

wing. As with the normal case, the narrowness of the resonance implies that the Boltzmann

exponential does not change much where the Lorentzian has a significant contribution, and so

we again will approximate e−E/T ≈ e−E/T . Since we avoid the resonant peak, the choice of E

not as straightforward in the subthreshold case where we took E = Eres. This choice makes

no sense in the subthreshold case, because the e−Eres/T > 1 in the subthreshold case, yet

obviously kinetic energy E > 0 and thus the Boltzmann factor must always be a suppression

144

and not an enhancement!

Yet clearly |Eres| remains an important scale. Thus we put E = u|Eres|, and we have

examined results for different values of the dimensionless parameter u. We find good agree-

ment with numerical results when we adopt u ≈ 1, i.e., E = |Eres|. Thus for the subthreshold

case we adopt a reaction rate which is in closely analogous to the normal case:

〈σv〉narrow, subthreshold = ω Γeff

(

µT

)3/2

e−|Eres|/Tf (−2|Eres|/Γtot) (A.10)

Similarly to the normal case, as the reaction becomes increasingly off-resonance, i.e., as

|Eres| grows, there is an exponential suppression. In addition, the correction factor has

limits f → 1/2 for |Eres| ≪ Γtot, and f → 0 as |Eres| ≫ Γtot. Finally, note that, as a function

of Eres, our subthreshold and normal rates match at Eres = 0, as they must physically.

145

Appendix B

B.1 The energy loss rates

The energy losses other than IC are bremsstrahlung, synchrotron and ionisation. They are

expressed as follows (Hayakawa, 1973; Ginzburg, 1979): synchrotron losses

bsync(Ee) =4

3σT c Umag

(

Ee

m c2

)2

=1

6 πσT cB2

(

Ee

m

)2

≈ 3 × 10−10 GeV sec−1

(

B

1 µG

)2 (Ee

10 TeV

)2

(B.1)

are controlled by the interstellar magnetic energy density Umag. The bremsstrahlung losses

depend on whether the medium is ionised or neutral. Due to the presence of both H I and

H II, the bremsstrahlung losses are due to both neutral and ionised hydrogen (Hayakawa,

1973; Ginzburg, 1979):

bbrem(Ee) = bbrem,ion(Ee) + bbrem,n(Ee) (B.2)

bbrem,ion(Ee) =3

πnH II α σT c Ee

[

ln

(

2 Ee

m c2

)

− 1

3

]

(B.3)

= 1.37 × 10−12 GeV sec−1( nH II

1 cm−3

)

×(

Ee

10 TeV

)[

ln

(

Ee

10 TeV

)

+ 17.2

]

(B.4)

146

bbrem,n(Ee) =3

πnH I α σT c Ee

[

ln(191) +1

18

]

(B.5)

= 7.3 × 10−12 GeV sec−1( nH I

1 cm−3

)

(

Ee

10 TeV

)

(B.6)

Here nH I and nH II are the number density of relativistic electrons in the interstellar medium,

which is equal to the number density of H I and H II.

147

References

Abazajian, K. N., Blanchet, S., & Harding, J. P. 2011, Physical Review D, 84, 103007

Abdo, A. A., Ackermann, M., Ajello, M., et al. 2009, Physical Review Letters, 103, 251101

—. 2010a, ApJ, 709, L152

—. 2010b, ApJ, 709, L152

—. 2010c, A&A, 523, A46

—. 2010d, A&A, 523, A46

—. 2010e, The Astrophysical Journal, 723, 1082

—. 2010f, A&A, 523, L2

—. 2010g, A&A, 523, L2

—. 2010h, A&A, 512, A7

—. 2010i, A&A, 512, A7

—. 2010j, Physical Review Letters, 104, 101101

—. 2010k, Physical Review Letters, 104, 101101

—. 2010l, The Astrophysical Journal, 720, 435

Abdo, A. A., Ackermann, M., Agudo, I., et al. 2010m, The Astrophysical Journal, 716, 30

Acero, F., Aharonian, F., Akhperjanian, A. G., et al. 2009, Science, 326, 1080

Ackermann, M., Ajello, M., Albert, A., et al. 2012a, Physical Review D, 85, 083007

—. 2012b, Physical Review D, 86, 022002

Ackermann, M., Ajello, M., Allafort, A., et al. 2012c, The Astrophysical Journal, 755, 164

Adahchour, A., & Descouvemont, P. 2003, Journal of Physics G Nuclear Physics, 29, 395

Agudo, I., Thum, C., Wiesemeyer, H., & Krichbaum, T. P. 2010, ApJS, 189, 1

148

Ahlers, M., Mertsch, P., & Sarkar, S. 2009, Physical Review D, 80, 123017

Ajzenberg-Selove, F. 1990, Nuclear Physics A, 506, 1

Ando, S., Cyburt, R. H., Hong, S. W., & Hyun, C. H. 2006, Physical Review C, 74, 025809

Ando, S., & Komatsu, E. 2006, Physical Review D, 73, 023521

Ando, S., Komatsu, E., Narumoto, T., & Totani, T. 2007, Physical Review D, 75, 063519

Ando, S., & Pavlidou, V. 2009, Mon. Not. Roy. Astron. Soc., 400, 2122

Angulo, C., Arnould, M., Rayet, M., et al. 1999, Nuclear Physics A, 656, 3

Angulo, C., Casarejos, E., Couder, M., et al. 2005, ApJ, 630, L105

Asplund, M., Lambert, D. L., Nissen, P. E., Primas, F., & Smith, V. V. 2006, The Astro-physical Journal, 644, 229

Baklanov, P. V., Blinnikov, S. I., & Pavlyuk, N. N. 2005, Astronomy Letters, 31, 429

Bania, T. M., Rood, R. T., & Balser, D. S. 2002, Nature, 415, 54

Bazin, G., Palanque-Delabrouille, N., Rich, J., et al. 2009, A&A, 499, 653

Becker, J. K. 2008, Phys. Rep., 458, 173

Berger, E., Kulkarni, S. R., Frail, D. A., & Soderberg, A. M. 2003, The AstrophysicalJournal, 599, 408

Bignami, G. F., Fichtel, C. E., Hartman, R. C., & Thompson, D. J. 1979, The AstrophysicalJournal, 232, 649

Blinov, D., Myserlis, I., Angelakis, E., King, O., & Pavlidou, V. 2013, The Astronomer’sTelegram, 4779, 1

Blumenthal, G. R., & Gould, R. J. 1970, Reviews of Modern Physics, 42, 237

Bonifacio, P. 2002, A&A, 395, 515

Bonifacio, P., Pasquini, L., Spite, F., et al. 2002, A&A, 390, 91

Bonometto, S., Cazzola, P., & Saggion, A. 1970, A&A, 7, 292

Botticella, M. T., Riello, M., Cappellaro, E., et al. 2008, A&A, 479, 49

Boyd, R. N., Brune, C. R., Fuller, G. M., & Smith, C. J. 2010a, Physical Review D, 82,105005

—. 2010b, Physical Review D, 82, 105005

Brecher, K., & Morrison, P. 1967, ApJ, 150, L61

149

Bringmann, T., Donato, F., & Lineros, R. A. 2012, JCAP, 1, 49

Cappellaro, E., Evans, R., & Turatto, M. 1999, A&A, 351, 459

Cappellaro, E., Riello, M., Altavilla, G., et al. 2005, A&A, 430, 83

Caraveo, P. A., Bennett, K., Bignami, G. F., et al. 1980, A&A, 91, L3

Carilli, C. L., & Rawlings, S. 2004, New Astronomy Review, 48, 979

Caughlan, G. R., & Fowler, W. A. 1988, Atomic Data and Nuclear Data Tables, 40, 283

Chakraborty, N., & Fields, B. D. 2012, ArXiv e-prints

Chakraborty, N., Fields, B. D., & Olive, K. A. 2011, Physical Review D, 83, 063006

Chang, Z., Jiang, Y., & Lin, H. N. 2013, ArXiv e-prints

Charity, R. J., Elson, J. M., Manfredi, J., et al. 2011, Physical Review C, 84, 014320

Chevalier, R. A. 1982a, ApJ, 259, L85

—. 1982b, The Astrophysical Journal, 259, 302

—. 1998, The Astrophysical Journal, 499, 810

Chomiuk, L., Soderberg, A. M., Chevalier, R., Badenes, C., & Fransson, C. 2011, in Bulletinof the American Astronomical Society, Vol. 43, American Astronomical Society MeetingAbstracts #217, #304.05

Cirelli, M., & Panci, P. 2009, Nuclear Physics B, 821, 399

Coc, A. 2013, Journal of Physics Conference Series, 420, 012136

Coc, A., Nunes, N. J., Olive, K. A., Uzan, J.-P., & Vangioni, E. 2007, Physical Review D,76, 023511

Coc, A., Vangioni-Flam, E., Descouvemont, P., Adahchour, A., & Angulo, C. 2004a, TheAstrophysical Journal, 600, 544

—. 2004b, The Astrophysical Journal, 600, 544

Coleman, S., & Glashow, S. L. 1999, Physical Review D, 59, 116008

Crockett, R. M., Smartt, S. J., Eldridge, J. J., et al. 2007, Mon. Not. Roy. Astron. Soc., 381,835

Cuoco, A., Iocco, F., Mangano, G., et al. 2004, International Journal of Modern Physics A,19, 4431

Cyburt, R. H. 2004a, Physical Review D, 70, 023505

150

—. 2004b, Physical Review D, 70, 023505

Cyburt, R. H., & Davids, B. 2008, Physical Review C, 78, 064614

Cyburt, R. H., Ellis, J., Fields, B. D., et al. 2010, JCAP, 10, 32

Cyburt, R. H., Ellis, J., Fields, B. D., Olive, K. A., & Spanos, V. C. 2006, JCAP, 11, 14

Cyburt, R. H., Fields, B. D., & Olive, K. A. 2001, New Astronomy, 6, 215

—. 2003, Physics Letters B, 567, 227

—. 2004, Physical Review D, 69, 123519

—. 2008, JCAP, 11, 12

Cyburt, R. H., Fields, B. D., Olive, K. A., & Skillman, E. 2005, Astroparticle Physics, 23,313

Cyburt, R. H., & Pospelov, M. 2012, International Journal of Modern Physics E, 21, 50004

Dahlen, T., Strolger, L., & Riess, A. G. 2010, in Bulletin of the American AstronomicalSociety, Vol. 42, American Astronomical Society Meeting Abstracts #215, #430.23

Dahlen, T., Strolger, L.-G., & Riess, A. G. 2008, The Astrophysical Journal, 681, 462

Dahlen, T., Strolger, L.-G., Riess, A. G., et al. 2004, The Astrophysical Journal, 613, 189

de Blok, E. W. J. G., Booth, R., Jonas, J., & Fanaroff, B. 2010, in ISKAF2010 ScienceMeeting

de Cea del Pozo, E., Torres, D. F., & Rodriguez Marrero, A. Y. 2009, The AstrophysicalJournal, 698, 1054

Dean, A. J., Clark, D. J., Stephen, J. B., et al. 2008, Science, 321, 1183

Dermer, C. D., Ramirez-Ruiz, E., & Le, T. 2007, ApJ, 664, L67

Dilday, B., Smith, M., Bassett, B., et al. 2010, The Astrophysical Journal, 713, 1026

Dmitriev, V. F., Flambaum, V. V., & Webb, J. K. 2004, Physical Review D, 69, 063506

Doert, M., Becker, J. K., Halzen, F., O’Murchadha, A., & Rhode, W. 2012, Journal ofPhysics Conference Series, 355, 012039

Eck, C. R., Cowan, J. J., Roberts, D. A., Boffi, F. R., & Branch, D. 1995, ApJ, 451, L53

Ellis, J., Farakos, K., Mavromatos, N. E., Mitsou, V. A., & Nanopoulos, D. V. 2000, TheAstrophysical Journal, 535, 139

Ellis, J., Olive, K. A., & Vangioni, E. 2005, Physics Letters B, 619, 30

151

Esmailzadeh, R., Starkman, G. D., & Dimopoulos, S. 1991, The Astrophysical Journal, 378,504

Faucher-Giguere, C.-A., & Loeb, A. 2010, JCAP, 1, 5

Felten, J. E. 1965, Physical Review Letters, 15, 1003

Felten, J. E., & Morrison, P. 1963, Physical Review Letters, 10, 453

Fermi-LAT Collaboration. 2013, ArXiv e-prints

Fichtel, C. E., Hartman, R. C., Kniffen, D. A., et al. 1975, The Astrophysical Journal, 198,163

Fields, B. D., Pavlidou, V., & Prodanovic, T. 2010, ApJ, 722, L199

Frieman, J. A., Bassett, B., Becker, A., et al. 2008, The Astronomical Journal, 135, 338

Fryer, C. L. 1999, The Astrophysical Journal, 522, 413

Fujiwara, M., Matsuoka, Y., & Ienaka, N. 2012, The Astronomical Journal, 144, 112

Fumagalli, M., O’Meara, J. M., & Prochaska, J. X. 2011, Science, 334, 1245

Gal-Yam, A., Fox, D. B., Price, P. A., et al. 2006, Nature, 444, 1053

Galama, T. J., Vreeswijk, P. M., van Paradijs, J., et al. 1998, Nature, 395, 670

Gehrels, N., Ramirez-Ruiz, E., & Fox, D. B. 2009, ARA&A, 47, 567

Geringer-Sameth, A., & Koushiappas, S. M. 2012, Mon. Not. Roy. Astron. Soc., 421, 1813

Ginzburg, V. L. 1979, Theoretical physics and astrophysics (Oxford: Pergamon)

Gonzalez Hernandez, J. I., Bonifacio, P., Caffau, E., et al. 2009, A&A, 505, L13

Gotz, D., Laurent, P., Lebrun, F., Daigne, F., & Bosnjak, Z. 2009, ApJ, 695, L208

Grindlay, J. E. 1978, Nature, 273, 211

H. E. S. S. Collaboration, Abramowski, A., Acero, F., et al. 2012, ArXiv e-prints

Hayakawa, S. 1973, Cosmic ray physics: nuclear and astrophysical aspects. (Moskva:Mir)

Heger, A., Fryer, C. L., Woosley, S. E., Langer, N., & Hartmann, D. H. 2003, The Astro-physical Journal, 591, 288

Heitler, W. 1954, Quantum theory of radiation

Helder, E. A., Vink, J., Bassa, C. G., et al. 2009, Science, 325, 719

Hensley, B. S., Siegal-Gaskins, J. M., & Pavlidou, V. 2010, The Astrophysical Journal, 723,277

152

Hopkins, A. M., & Beacom, J. F. 2006, The Astrophysical Journal, 651, 142

Horiuchi, S., Beacom, J. F., & Dwek, E. 2009a, Physical Review D, 79, 083013

—. 2009b, Physical Review D, 79, 083013

Horiuchi, S., Beacom, J. F., Kochanek, C. S., et al. 2011, The Astrophysical Journal, 738,154

Howk, J. C., Lehner, N., Fields, B. D., & Mathews, G. J. 2012, Nature, 489, 121

Hoyle, F. 1954, ApJS, 1, 121

Hunter, S. D., Bertsch, D. L., Catelli, J. R., et al. 1997, The Astrophysical Journal, 481, 205

Iben, Jr., I., & Tutukov, A. V. 1984, ApJS, 54, 335

Iwamoto, K., Mazzali, P. A., Nomoto, K., et al. 1998, Nature, 395, 672

Izotov, Y. I., Chaffee, F. H., Foltz, C. B., et al. 1999, The Astrophysical Journal, 527, 757

Jedamzik, K. 2004, Physical Review D, 70, 063524

Jedamzik, K., & Pospelov, M. 2009, New Journal of Physics, 11, 105028

Johnston, S., Feain, I. J., & Gupta, N. 2009, in Astronomical Society of the Pacific Confer-ence Series, Vol. 407, The Low-Frequency Radio Universe, ed. D. J. Saikia, D. A. Green,Y. Gupta, & T. Venturi, 446

Jones, F. C. 1968, Physical Review, 167, 1159

Jorstad, S. G., Marscher, A. P., Larionov, V. M., et al. 2010, The Astrophysical Journal,715, 362

Kalemci, E., Boggs, S., Wunderer, C., & Jean, P. 2004, in ESA Special Publication, Vol. 552,5th INTEGRAL Workshop on the INTEGRAL Universe, ed. V. Schoenfelder, G. Lichti,& C. Winkler, 859

Kazanas, D., & Protheroe, R. J. 1983, Nature, 302, 228

Kennicutt, Jr., R. C. 1998, The Astrophysical Journal, 498, 541

Kirsebom, O. S., & Davids, B. 2011, Physical Review C, 84, 058801

Kistler, M. D., Yuksel, H., Ando, S., Beacom, J. F., & Suzuki, Y. 2011, Physical Review D,83, 123008

Kneller, J. P., Scherrer, R. J., Steigman, G., & Walker, T. P. 2001, Physical Review D, 64,123506

Kolb, E. W., & Turner, M. S. 1990, The early universe.

153

Komatsu, E., Dunkley, J., Nolta, M. R., et al. 2009, ApJS, 180, 330

Komatsu, E., Smith, K. M., Dunkley, J., et al. 2011, ApJS, 192, 18

Korn, A. J., Grundahl, F., Richard, O., et al. 2006, Nature, 442, 657

Kraushaar, W. L., Clark, G. W., Garmire, G. P., et al. 1972, The Astrophysical Journal,177, 341

Krawczynski, H. 2012, The Astrophysical Journal, 744, 30

Krawczynski, H., Angelini, L., Baring, M., et al. 2013, ArXiv e-prints

Kudritzki, R.-P., & Puls, J. 2000, ARA&A, 38, 613

Kulkarni, S. R., Frail, D. A., Wieringa, M. H., et al. 1998, Nature, 395, 663

Lacki, B. C., Horiuchi, S., & Beacom, J. F. 2012, ArXiv e-prints

Lacki, B. C., Thompson, T. A., & Quataert, E. 2010, The Astrophysical Journal, 717, 1

Lacki, B. C., Thompson, T. A., Quataert, E., Loeb, A., & Waxman, E. 2011, The Astro-physical Journal, 734, 107

Laurent, P., Gotz, D., Binetruy, P., Covino, S., & Fernandez-Soto, A. 2011, Physical ReviewD, 83, 121301

Lazzati, D. 2006, New Journal of Physics, 8, 131

Lebrun, F., Paul, J. A., Bignami, G. F., et al. 1982, A&A, 107, 390

Leitherer, C., Robert, C., & Drissen, L. 1992, The Astrophysical Journal, 401, 596

Lenain, J.-P., & Walter, R. 2011, A&A, 535, A19

Li, W., Chornock, R., Leaman, J., et al. 2011a, Mon. Not. Roy. Astron. Soc., 412, 1473

—. 2011b, Mon. Not. Roy. Astron. Soc., 412, 1473

Li, W., Leaman, J., Chornock, R., et al. 2011c, Mon. Not. Roy. Astron. Soc., 412, 1441

Li, Z.-Y., & Chevalier, R. A. 1999, The Astrophysical Journal, 526, 716

Lien, A., Chakraborty, N., Fields, B. D., & Kemball, A. 2011, The Astrophysical Journal,740, 23

Lien, A., & Fields, B. D. 2009, JCAP, 1, 47

—. 2012, ArXiv e-prints

Lien, A., Fields, B. D., & Beacom, J. F. 2010, Physical Review D, 81, 083001

Loeb, A., & Waxman, E. 2000, Nature, 405, 156

154

LSST Science Collaboration, Abell, P. A., Allison, J., et al. 2009, ArXiv e-prints

MacFadyen, A. I., & Woosley, S. E. 1999, The Astrophysical Journal, 524, 262

MacFadyen, A. I., Woosley, S. E., & Heger, A. 2001, The Astrophysical Journal, 550, 410

Madau, P., della Valle, M., & Panagia, N. 1998, Mon. Not. Roy. Astron. Soc., 297, L17

Makiya, R., Totani, T., & Kobayashi, M. A. R. 2011, The Astrophysical Journal, 728, 158

Malyshev, D., & Hogg, D. W. 2011, The Astrophysical Journal, 738, 181

Mannucci, F., Della Valle, M., & Panagia, N. 2007, Mon. Not. Roy. Astron. Soc., 377, 1229

Mayer-Hasselwander, H. A., Buccheri, R., Kanbach, G., et al. 1980, in Annals of the NewYork Academy of Sciences, Vol. 336, Ninth Texas Symposium on Relativistic Astrophysics,ed. J. Ehlers, J. J. Perry, & M. Walker, 211–222

McConnell, M. L., Smith, D. M., Emslie, A. G., et al. 2003, in Bulletin of the AmericanAstronomical Society, Vol. 35, AAS/Solar Physics Division Meeting #34, 850

McConville, W., Ostorero, L., Moderski, R., et al. 2011, The Astrophysical Journal, 738, 148

McNamara, A. L., Kuncic, Z., & Wu, K. 2009, Mon. Not. Roy. Astron. Soc., 395, 1507

Melendez, J., Casagrande, L., Ramırez, I., Asplund, M., & Schuster, W. J. 2010a, A&A,515, L3

Melendez, J., Ramırez, I., Casagrande, L., et al. 2010b, Astrophysics and Space Science, 328,193

Meyer, E. T., Fossati, G., Georganopoulos, M., & Lister, M. L. 2012, ApJ, 752, L4

Miniati, F. 2002, Mon. Not. Roy. Astron. Soc., 337, 199

Mokiem, M. R., de Koter, A., Vink, J. S., et al. 2007, A&A, 473, 603

Montes, M. J., Weiler, K. W., & Panagia, N. 1997, The Astrophysical Journal, 488, 792

Moskalenko, I. V., & Porter, T. A. 2009, ApJ, 692, L54

Moskalenko, I. V., & Strong, A. W. 2000, The Astrophysical Journal, 528, 357

Mucke, A., & Pohl, M. 2000, Mon. Not. Roy. Astron. Soc., 312, 177

Mukhanov, V. 2004, International Journal of Theoretical Physics, 43, 669

Munari, U., Barbon, R., Piemonte, A., Tomasella, L., & Rejkuba, M. 1998, A&A, 333, 159

Murphy, E. J., Porter, T. A., Moskalenko, I. V., Helou, G., & Strong, A. W. 2012, TheAstrophysical Journal, 750, 126

155

Nakano, S., & Aoki, M. 1997, IAU circ., 6795, 2

Nalewajko, K. 2013, Mon. Not. Roy. Astron. Soc., 598

Nolan, P. L., Abdo, A. A., Ackermann, M., et al. 2012a, ApJS, 199, 31

—. 2012b, ApJS, 199, 31

Nomoto, K., Thielemann, F.-K., & Yokoi, K. 1984, The Astrophysical Journal, 286, 644

Novick, R. 1975, Space Science Reviews, 18, 389

O’Malley, P. D., Bardayan, D. W., Adekola, A. S., et al. 2011, Physical Review C, 84, 042801

O’Meara, J. M., Burles, S., Prochaska, J. X., et al. 2006, ApJ, 649, L61

Padovani, P., Ghisellini, G., Fabian, A. C., & Celotti, A. 1993, Mon. Not. Roy. Astron. Soc.,260, L21

Paglione, T. A. D., & Abrahams, R. D. 2012, The Astrophysical Journal, 755, 106

Palanque-Delabrouille, N., Ruhlmann-Kleider, V., Pascal, S., et al. 2010, A&A, 514, A63

Panagia, N., Van Dyk, S. D., Weiler, K. W., et al. 2006, The Astrophysical Journal, 646,369

Papenkova, M., Li, W. D., Wray, J., Chleborad, C. W., & Schwartz, M. 2001, IAU circ.,7722, 1

Pavlidou, V., & Fields, B. D. 2001, The Astrophysical Journal, 558, 63

—. 2002, ApJ, 575, L5

Persic, M., & Rephaeli, Y. 2010, Mon. Not. Roy. Astron. Soc., 403, 1569

—. 2011, ArXiv e-prints

—. 2012, ArXiv e-prints

Phillips, M. M. 1993, ApJ, 413, L105

Pieper, S. C., Varga, K., & Wiringa, R. B. 2002, Physical Review C, 66, 044310

Pohl, M. 1993, A&A, 270, 91

—. 1994, A&A, 287, 453

Pooley, D., Lewin, W. H. G., Fox, D. W., et al. 2002, The Astrophysical Journal, 572, 932

Porter, T. A., Moskalenko, I. V., Strong, A. W., Orlando, E., & Bouchet, L. 2008, TheAstrophysical Journal, 682, 400

Price, P. A., Berger, E., Reichart, D. E., et al. 2002, ApJ, 572, L51

156

Regis, M., & Ullio, P. 2008, Physical Review D, 78, 043505

Reimer, A. 2012, Journal of Physics Conference Series, 355, 012011

Riess, A. G., Press, W. H., & Kirshner, R. P. 1996, The Astrophysical Journal, 473, 88

Robitaille, T. P., & Whitney, B. A. 2010, ApJ, 710, L11

Rossi, E. M., Lazzati, D., Salmonson, J. D., & Ghisellini, G. 2004, Mon. Not. Roy. Astron.Soc., 354, 86

Ryan, S. G., Beers, T. C., Olive, K. A., Fields, B. D., & Norris, J. E. 2000, ApJ, 530, L57

Rybicki, G. B., & Lightman, A. P. 1986, Radiative Processes in Astrophysics

Sako, M., Bassett, B., Becker, A., et al. 2008, The Astronomical Journal, 135, 348

Salamon, M. H., & Stecker, F. W. 1998, The Astrophysical Journal, 493, 547

Salpeter, E. E. 1955, The Astrophysical Journal, 121, 161

Sanders, D. B., Mazzarella, J. M., Kim, D.-C., Surace, J. A., & Soifer, B. T. 2003, TheAstronomical Journal, 126, 1607

Scheuer, P. A. G. 1957, Proceedings of the Cambridge Philosophical Society, 53, 764

Smartt, S. J., Eldridge, J. J., Crockett, R. M., & Maund, J. R. 2009, Mon. Not. Roy. Astron.Soc., 395, 1409

Smith, M. S., Kawano, L. H., & Malaney, R. A. 1993, ApJS, 85, 219

Soderberg, A. M. 2007, in American Institute of Physics Conference Series, Vol. 937, Su-pernova 1987A: 20 Years After: Supernovae and Gamma-Ray Bursters, ed. S. Immler,K. Weiler, & R. McCray, 492–499

Soderberg, A. M., Chevalier, R. A., Kulkarni, S. R., & Frail, D. A. 2006a, The AstrophysicalJournal, 651, 1005

Soderberg, A. M., Frail, D. A., & Wieringa, M. H. 2004, ApJ, 607, L13

Soderberg, A. M., Nakar, E., Berger, E., & Kulkarni, S. R. 2006b, The Astrophysical Journal,638, 930

Sorcia, M., Benıtez, E., Hiriart, D., et al. 2013, ArXiv e-prints

Spite, F., & Spite, M. 1982, A&A, 115, 357

Sreekumar, P., Bertsch, D. L., Dingus, B. L., et al. 1998, The Astrophysical Journal, 494,523

Stecker, F. W., Malkan, M. A., & Scully, S. T. 2012, ArXiv e-prints

157

Stecker, F. W., Salamon, M. H., & Malkan, M. A. 1993, ApJ, 410, L71

Stecker, F. W., & Venters, T. M. 2011, The Astrophysical Journal, 736, 40

Steigman, G. 2007, Annual Review of Nuclear and Particle Science, 57, 463

Stockdale, C. J., Kelley, M. T., Weiler, K. W., et al. 2007, in American Institute of PhysicsConference Series, Vol. 937, Supernova 1987A: 20 Years After: Supernovae and Gamma-Ray Bursters, ed. S. Immler, K. Weiler, & R. McCray, 264–268

Stockdale, C. J., Weiler, K. W., Van Dyk, S. D., et al. 2003, The Astrophysical Journal, 592,900

Strong, A. W., & Moskalenko, I. V. 1998, The Astrophysical Journal, 509, 212

Strong, A. W., Moskalenko, I. V., & Reimer, O. 2000, The Astrophysical Journal, 537, 763

Strong, A. W., Orlando, E., & Jaffe, T. R. 2011, A&A, 534, A54

Strong, A. W., Porter, T. A., Digel, S. W., et al. 2010, ApJ, 722, L58

Strong, A. W., Wolfendale, A. W., & Worrall, D. M. 1976a, Journal of Physics A Mathe-matical General, 9, 1553

—. 1976b, Mon. Not. Roy. Astron. Soc., 175, 23P

—. 1976c, Mon. Not. Roy. Astron. Soc., 175, 23P

Teichmann, T., & Wigner, E. P. 1952, Physical Review, 87, 123

Thompson, T. A., Quataert, E., & Waxman, E. 2007, The Astrophysical Journal, 654, 219

Tilley, D. R., Kelley, J. H., Godwin, J. L., et al. 2004, Nuclear Physics A, 745, 155

Torres, D. F., Reimer, O., Domingo-Santamarıa, E., & Digel, S. W. 2004, ApJ, 607, L99

Uchiyama, Y., Aharonian, F. A., Tanaka, T., Takahashi, T., & Maeda, Y. 2007, Nature,449, 576

VERITAS Collaboration, Acciari, V. A., Aliu, E., et al. 2009, Nature, 462, 770

Vink, J. S., de Koter, A., & Lamers, H. J. G. L. M. 2001, A&A, 369, 574

Voelk, H. J. 1989, A&A, 218, 67

Volk, H. J., Aharonian, F. A., & Breitschwerdt, D. 1996, Space Science Reviews, 75, 279

Wagoner, R. V., Fowler, W. A., & Hoyle, F. 1967a, The Astrophysical Journal, 148, 3

—. 1967b, The Astrophysical Journal, 148, 3

158

Wainscoat, R. J., Cohen, M., Volk, K., Walker, H. J., & Schwartz, D. E. 1992, ApJS, 83,111

Wardle, J. F. C., & Kronberg, P. P. 1974, The Astrophysical Journal, 194, 249

Webber, W. R., Simpson, G. A., & Cane, H. V. 1980, The Astrophysical Journal, 236, 448

Webbink, R. F. 1984, The Astrophysical Journal, 277, 355

Wehrle, A. E., Marscher, A. P., Jorstad, S. G., et al. 2012, The Astrophysical Journal, 758,72

Weiler, K. W., Panagia, N., Montes, M. J., & Sramek, R. A. 2002, ARA&A, 40, 387

Weiler, K. W., Panagia, N., & Sramek, R. A. 1990, The Astrophysical Journal, 364, 611

Weiler, K. W., Panagia, N., Sramek, R. A., et al. 2009, in American Institute of Physics Con-ference Series, Vol. 1111, American Institute of Physics Conference Series, ed. G. Giobbi,A. Tornambe, G. Raimondo, M. Limongi, L. A. Antonelli, N. Menci, & E. Brocato, 440–447

Weiler, K. W., Sramek, R. A., Panagia, N., van der Hulst, J. M., & Salvati, M. 1986, TheAstrophysical Journal, 301, 790

Weiler, K. W., van Dyk, S. D., Sramek, R. A., & Panagia, N. 2004, New Astronomy Review,48, 1377

Weiler, K. W., Williams, C. L., Panagia, N., et al. 2007, The Astrophysical Journal, 671,1959

Weiß, A., Neininger, N., Huttemeister, S., & Klein, U. 2001, A&A, 365, 571

Weisskopf, M. C., Elsner, R. F., & O’Dell, S. L. 2010, in Society of Photo-Optical In-strumentation Engineers (SPIE) Conference Series, Vol. 7732, Society of Photo-OpticalInstrumentation Engineers (SPIE) Conference Series

Wellons, S., & Soderberg, A. M. 2011, in Bulletin of the American Astronomical Society,Vol. 43, American Astronomical Society Meeting Abstracts #217, #337.15

Woosley, S. E. 1993, The Astrophysical Journal, 405, 273

Woosley, S. E., & Bloom, J. S. 2006, ARA&A, 44, 507

Woosley, S. E., Eastman, R. G., & Schmidt, B. P. 1999, The Astrophysical Journal, 516,788

Yan, J., Cecil, F. E., Greife, U., et al. 2002, Physical Review C, 65, 048801

Zacharias, M., & Schlickeiser, R. 2012, Mon. Not. Roy. Astron. Soc., 420, 84

159